Download as pdf or txt
Download as pdf or txt
You are on page 1of 69

Philosophy and Model Theory Tim

Button
Visit to download the full and correct content document:
https://ebookmass.com/product/philosophy-and-model-theory-tim-button/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Model Predictive Control: Theory, Computation, and


Design, 2nd Edition Rawlings James B.

https://ebookmass.com/product/model-predictive-control-theory-
computation-and-design-2nd-edition-rawlings-james-b/

Philosophy of Film Without Theory Craig Fox

https://ebookmass.com/product/philosophy-of-film-without-theory-
craig-fox/

Gender Issues and Philosophy Education: History –


Theory – Practice Markus Tiedemann

https://ebookmass.com/product/gender-issues-and-philosophy-
education-history-theory-practice-markus-tiedemann/

The Art of Anatheism (Reframing Continental Philosophy


of Religion) 1st Edition Tim Di Muzio

https://ebookmass.com/product/the-art-of-anatheism-reframing-
continental-philosophy-of-religion-1st-edition-tim-di-muzio/
The Philosophy of Evolutionary Theory: Concepts,
Inferences, and Probabilities 1st Edition Elliott Sober

https://ebookmass.com/product/the-philosophy-of-evolutionary-
theory-concepts-inferences-and-probabilities-1st-edition-elliott-
sober/

The Politics of Vulnerable Groups: Implications for


Philosophy, Law, and Political Theory Fabio Macioce

https://ebookmass.com/product/the-politics-of-vulnerable-groups-
implications-for-philosophy-law-and-political-theory-fabio-
macioce/

Indonesian Law Tim Lindsey

https://ebookmass.com/product/indonesian-law-tim-lindsey/

Boxing the Octopus Tim Maleeny

https://ebookmass.com/product/boxing-the-octopus-tim-maleeny/

Landscape, Materiality and Heritage: An Object


Biography Tim Edensor

https://ebookmass.com/product/landscape-materiality-and-heritage-
an-object-biography-tim-edensor/
Philosophy and Model Theory
Philosophy and
Model Theory

Tim Button and Sean Walsh


with a historical appendix by Wilfrid Hodges
3
Great Clarendon Street, Oxford, ox2 6dp,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Tim Button and Sean Walsh 2018
© Historical Appendix D Wilfrid Hodges
The moral rights of the authors have been asserted
First Edition published in 2018
Impression: 1
All rights reserved. No part of this publication may be reproduced, stored in
a retrieval system, or transmitted, in any form or by any means, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence, or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
You must not circulate this work in any other form
and you must impose this same condition on any acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
Library of Congress Control Number: 2017959066
ISBN: 978–0–19–879039–6 (hbk.)
978–0–19–879040–2 (pbk.)
Printed and bound by
CPI Group (UK) Ltd, Croydon, cr0 4yy
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
Preface

Philosophy and model theory frequently meet one another. This book aims to un-
derstand their interactions.
Model theory is used in every ‘theoretical’ branch of analytic philosophy: in phi-
losophy of mathematics; in philosophy of science; in philosophy of language; in
philosophical logic; and in metaphysics. But these wide-ranging appeals to model
theory have created a highly fragmented literature. On the one hand, many philo-
sophically significant results are found only in mathematics textbooks: these are
aimed squarely at mathematicians; they typically presuppose that the reader has a
serious background in mathematics; and little clue is given as to their philosophical
significance. On the other hand, the philosophical applications of these results are
scattered across disconnected pockets of papers.
The first aim of our book, then, is to consider the philosophical uses of model the-
ory. We state and prove the best versions of results for philosophical purposes. We
then probe their philosophical significance. And we show how similar dialectical
situations arise repeatedly across fragmented debates in different areas.
The second aim of our book, though, is to consider the philosophy of model theory.
Model theory itself is rarely taken as the subject matter of philosophising (contrast
this with the philosophy of biology, or the philosophy of set theory). But model
theory is a beautiful part of pure mathematics, and worthy of philosophical study
in its own right.
Both aims give rise to challenges. On the one hand: the philosophical uses of
model theory are scattered across a disunified literature. And on the other hand:
there is scarcely any literature on the philosophy of model theory.
All of which is to say: philosophy and model theory isn’t really ‘a thing’ yet. This
book aims to start carving out such a thing. We want to chart the rock-face and
trace its dialectical contours. But we present this book, not as a final word on what
philosophically inclined model theorists and model-theoretically inclined philoso-
phers should do, but as an invitation to join in.
So. This is not a book in which a single axe is ground, page by page, into an
increasingly sharp blade. No fundamental line of argument—arching from Chapter
1 through to Chapter 17—serves as the spine of the book. What knits the chapters
together into a single book is not a single thesis, but a sequence of overlapping,
criss-crossing themes.
vi preface

Topic selection
Precisely because philosophy and model theory isn’t yet ‘a thing’, we have had to
make some difficult decisions about what topics to discuss.
On the one hand, we aimed to pick topics which should be of fairly mainstream
philosophical concern. So, when it comes to the philosophical uses of model the-
ory, we have largely considered topics concerning reference, realism, and doxology
(a term we introduce in Chapter 6). But, even when we have considered questions
which fall squarely within in the philosophy of model theory, the questions that
we have focussed on are clearly instances of ‘big questions’. We look at questions
of sameness of theories/structure (Chapter 5); of taking diverse perspectives on the
same concept (in Chapter 14); of how to draw boundaries of logic (in Chapter 16);
and of classification of mathematical objects (Chapter 17).
On the other hand, we also wanted to give you a decent bang for your buck. We
figured that if you were going to have to wrestle with some new philosophical idea
or model-theoretic result, then you should get to see it put to decent use. (This ex-
plains why, for example, the Push-Through Construction, the just-more theory ma-
noeuvre, supervaluational semantics, and the ideas of moderation and modelism,
occur so often in this book.) Conversely, we have had to set aside debates—no
matter how interesting—which would have taken too long to set up.
All of which is to say: this book is not comprehensive. Not even close.
Although we consider models of set theory in Chapters 8 and 11, we scarcely
scratch the surface. Although we discuss infinitary logics in Chapters 15–16, we only
use them in fairly limited ways.1 Whilst we mention Tennenbaum’s Theorem in
Chapter 7, that is as close as we get to computable model theory. We devote only
one brief section to o-minimality, namely §4.10. And although we consider quanti-
fiers in Chapter 16 and frequently touch on issues concerning logical consequence,
we never address the latter topic head on.2
The grave enormity, though, is that we have barely scratched the surface of model
theory itself. As the table of contents reveals, the vast majority of the book considers
model theory as it existed before Morley’s Categoricity Theorem.
Partially correcting for this, Wilfrid Hodges’ wonderful historical essay appears
as Part D of this book. Wilfrid’s essay treats Morley’s Theorem as a pivot-point for
the subject of model theory. He looks back critically to the history, to uncover the
notions at work in Morley’s Theorem and its proof, and he looks forward to the
riches that have followed from it. We have both learned so much from Wilfrid’s
Model Theory,3 and we are delighted to include his ‘short history’ here.

1 We do not, for example, consider the connection between infinitary logics and supervenience, as

Glanzberg (2001) and Bader (2011) do.


2 For that, we would point the reader to Blanchette (2001) and Shapiro (2005a).
3 Hodges (1993).
preface vii

Still, concerning all those topics which we have omitted: we intend no slight
against them. But there is only so much one book can do, and this book is already
much (much) longer than we originally planned. We are sincere in our earlier claim,
that this book is not offered as a final word, but as an invitation to take part.

Structuring the book


Having selected our topics, we needed to arrange them into a book. At this point,
we realised that these topics have no natural linear ordering.
As such, we have tried to strike a balance between three aims that did not always
point in the same direction: to order by philosophical theme, to order by increasing
philosophical sophistication, and to order by increasing mathematical sophistication.
The book’s final structure of represents our best compromise between these three
aims. It is divided into three main parts: Reference and realism, Categoricity, and
Indiscernibility and classification.
Each part has an introduction, and those who want to dip in and out of particular
topics, rather than reading cover-to-cover, should read the three part-introductions
after they have finished reading this preface. The part-introductions provide the-
matic overviews of each chapter, and they also contain diagrams which depict the
dependencies between each section of the book. In combination with the table of
contents, these diagrams will allow readers to take shortcuts to their favourite des-
tinations, without having to stop to smell every rose along the way.

Presuppositions and proofs


So far as possible, the book assumes only that you have completed a 101-level logic
course, and so have some familiarity with first-order logic.
Inevitably, there are some exceptions to this: we were forced to assume some
familiarity with analysis when discussing infinitesimals in Chapter 4, and equally
some familiarity with topology when discussing Stone spaces in Chapter 14. We do
not prove Gödel’s incompleteness results, although we do state versions of them in
§5.a. A book can only be so self-contained.
By and large, though, this book is self-contained. When we invoke a model-
theoretic notion, we almost always define the notion formally in the text. When
it comes to proofs, we follow these rules of thumb.
The main text includes both brief proofs, and also those proofs which we wanted
to discuss directly.
The appendices include proofs which we wanted to include in the book, but
which were too long to feature in the main text. These include: proofs concern-
ing elementary topics which our readers should come to understand (at least one
viii preface

day); proofs which are difficult to access in the existing literature; proofs of certain
folk-lore results; and proofs of new results.
But the book omits all proofs which are both readily accessed and too long to be
self-contained. In such cases, we simply provide readers with citations.
The quick moral for readers to extract is this. If you encounter a proof in the main
text of a chapter, you should follow it through. But we would add a note for read-
ers whose primary background is in philosophy. If you really want to understand a
mathematical concept, you need to see it in action. Read the appendices!

Acknowledgements
The book arose from a seminar series on philosophy and model theory that we ran
in Birkbeck in Autumn 2011. We turned the seminar into a paper, but it was vastly
too long. An anonymous referee for Philosophia Mathematica suggested the paper
might form the basis for a book. So it did.
We have presented topics from this book several times. It would not be the
book it is, without the feedback, questions and comments we have received. So
we owe thanks to: an anonymous referee for Philosophia Mathematica, and James
Studd for OUP; and to Sarah Acton, George Anegg, Andrew Arana, Bahram Assa-
dian, John Baldwin, Kyle Banick, Neil Barton, Timothy Bays, Anna Bellomo, Liam
Bright, Chloé de Canson, Adam Caulton, Catrin Campbell-Moore, John Corco-
ran, Radin Dardashti, Walter Dean, Natalja Deng, William Demopoulos, Michael
Detlefsen, Fiona Doherty, Cian Dorr, Stephen Duxbury, Sean Ebels-Duggan, Sam
Eklund, Hartry Field, Branden Fitelson, Vera Flocke, Salvatore Florio, Peter Fritz,
Michael Gabbay, Haim Gaifman, J. Ethan Galebach, Marcus Giaquinto, Peter Gib-
son, Tamara von Glehn, Owen Griffiths, Emmylou Haffner, Bob Hale, Jeremy
Heis, Will Hendy, Simon Hewitt, Kate Hodesdon, Wilfrid Hodges, Luca Incur-
vati, Douglas Jesseph, Nicholas Jones, Peter Koellner, Brian King, Eleanor Knox, Jo-
hannes Korbmacher, Arnold Koslow, Hans-Christoph Kotzsch, Greg Lauro, Sarah
Lawsky, Øystein Linnebo, Yang Liu, Pen Maddy, Kate Manion, Tony Martin, Guil-
laume Massas, Vann McGee, Toby Meadows, Richard Mendelsohn, Christopher
Mitsch, Stella Moon, Adrian Moore, J. Brian Pitts, Jonathan Nassim, Fredrik Ny-
seth, Sara Parhizgari, Charles Parsons, Jonathan Payne, Graham Priest, Michael
Potter, Hilary Putnam, Paula Quinon, David Rabouin, Erich Reck, Sam Roberts,
Marcus Rossberg, J. Schatz, Gil Sagi, Bernhard Salow, Chris Scambler, Thomas
Schindler, Dana Scott, Stewart Shapiro, Gila Sher, Lukas Skiba, Jönne Speck, Se-
bastian Speitel, Will Stafford, Trevor Teitel, Robert Trueman, Jouko Väänänen, Kai
Wehmeier, J. Robert G. Williams, John Wigglesworth, Hugh Woodin, Jack Woods,
Crispin Wright, Wesley Wrigley, and Kino Zhao.
We owe some special debts to people involved in the original Birkbeck seminar.
preface ix

First, the seminar was held under the auspices of the Department of Philosophy at
Birkbeck and Øystein Linnebo’s European Research Council-funded project ‘Plu-
rals, Predicates, and Paradox’, and we are very grateful to all the people from the
project and the department for participating and helping to make the seminar pos-
sible. Second, we were lucky to have several great external speakers visit the sem-
inar, whom we would especially like to thank. The speakers were: Timothy Bays,
Walter Dean, Volker Halbach, Leon Horsten, Richard Kaye, Jeff Ketland, Angus
Macintyre, Paula Quinon, Peter Smith, and J. Robert G. Williams. Third, many of
the external talks were hosted by the Institute of Philosophy, and we wish to thank
Barry C. Smith and Shahrar Ali for all their support and help in this connection.
A more distant yet important debt is owed to Denis Bonnay, Brice Halimi, and
Jean-Michel Salanskis, who organised a lovely event in Paris in June 2010 called ‘Phi-
losophy and Model Theory.’ That event got some of us first thinking about ‘Philos-
ophy and Model Theory’ as a unified topic.
We are also grateful to various editors and publishers for allowing us to reuse pre-
viously published material. Chapter 5 draws heavily on Walsh 2014, and the copy-
right is held by the Association for Symbolic Logic and is being used with their
permission. Chapters 7–11 draw heavily upon on Button and Walsh 2016, published
by Philosophia Mathematica. Finally, §13.7 draws from Button 2016b, published by
Analysis, and §15.1 draws from Button 2017, published by the Notre Dame Journal of
Formal Logic.
Finally, though, a word from us, as individuals.
From Tim. I want to offer my deep thanks to the Leverhulme Trust: their fund-
ing, in the form of a Philip Leverhulme Prize (plp–2014–140), enabled me to take
the research leave necessary for this book. But I mostly want to thank two very spe-
cial people. Without Sean, this book could not be. And without my Ben, I could
not be.
From Sean. I want to thank the Kurt Gödel Society, whose funding, in the form
of a Kurt Gödel Research Prize Fellowship, helped us put on the original Birkbeck
seminar. I also want to thank Tim for being a model co-author and a model friend.
Finally, I want to thank Kari for her complete love and support.
Contents

A Reference and realism 1

1 Logics and languages 7


1.1 Signatures and structures . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 First-order logic: a first look . . . . . . . . . . . . . . . . . . . . . . 9
1.3 The Tarskian approach to semantics . . . . . . . . . . . . . . . . . 12
1.4 Semantics for variables . . . . . . . . . . . . . . . . . . . . . . . . 13
1.5 The Robinsonian approach to semantics . . . . . . . . . . . . . . 15
1.6 Straining the notion of ‘language’ . . . . . . . . . . . . . . . . . . . 17
1.7 The Hybrid approach to semantics . . . . . . . . . . . . . . . . . . 18
1.8 Linguistic compositionality . . . . . . . . . . . . . . . . . . . . . . 19
1.9 Second-order logic: syntax . . . . . . . . . . . . . . . . . . . . . . 21
1.10 Full semantics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.11 Henkin semantics . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.12 Consequence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.13 Definability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
1.a First- and second-order arithmetic . . . . . . . . . . . . . . . . . . 28
1.b First- and second-order set theory . . . . . . . . . . . . . . . . . . 30
1.c Deductive systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

2 Permutations and referential indeterminacy 35


2.1 Isomorphism and the Push-Through Construction . . . . . . . . . 35
2.2 Benacerraf ’s use of Push-Through . . . . . . . . . . . . . . . . . . 37
2.3 Putnam’s use of Push-Through . . . . . . . . . . . . . . . . . . . . 39
2.4 Attempts to secure reference in mathematics . . . . . . . . . . . . 44
2.5 Supervaluationism and indeterminacy . . . . . . . . . . . . . . . . 47
2.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.a Eligibility, definitions, and Completeness . . . . . . . . . . . . . . 50
2.b Isomorphism and satisfaction . . . . . . . . . . . . . . . . . . . . . 52

3 Ramsey sentences and Newman’s objection 55


3.1 The o/t dichotomy . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.2 Ramsey sentences . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.3 The promise of Ramsey sentences . . . . . . . . . . . . . . . . . . 57
3.4 A caveat on the o/t dichotomy . . . . . . . . . . . . . . . . . . . . 58
3.5 Newman’s criticism of Russell . . . . . . . . . . . . . . . . . . . . . 59
xii contents

3.6 The Newman-conservation-objection . . . . . . . . . . . . . . . . 60


3.7 Observation vocabulary versus observable objects . . . . . . . . . 63
3.8 The Newman-cardinality-objection . . . . . . . . . . . . . . . . . 64
3.9 Mixed-predicates again: the case of causation . . . . . . . . . . . . 66
3.10 Natural properties and just more theory . . . . . . . . . . . . . . . 67
3.a Newman and elementary extensions . . . . . . . . . . . . . . . . . 69
3.b Conservation in first-order theories . . . . . . . . . . . . . . . . . 72

4 Compactness, infinitesimals, and the reals 75


4.1 The Compactness Theorem . . . . . . . . . . . . . . . . . . . . . 75
4.2 Infinitesimals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.3 Notational conventions . . . . . . . . . . . . . . . . . . . . . . . . 79
4.4 Differentials, derivatives, and the use of infinitesimals . . . . . . . 79
4.5 The orders of infinite smallness . . . . . . . . . . . . . . . . . . . . 81
4.6 Non-standard analysis with a valuation . . . . . . . . . . . . . . . 84
4.7 Instrumentalism and conservation . . . . . . . . . . . . . . . . . . 88
4.8 Historical fidelity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.9 Axiomatising non-standard analysis . . . . . . . . . . . . . . . . . 93
4.10 Axiomatising the reals . . . . . . . . . . . . . . . . . . . . . . . . . 97
4.a Gödel’s Completeness Theorem . . . . . . . . . . . . . . . . . . . 99
4.b A model-theoretic proof of Compactness . . . . . . . . . . . . . . 103
4.c The valuation function of §4.6 . . . . . . . . . . . . . . . . . . . . 104

5 Sameness of structure and theory 107


5.1 Definitional equivalence . . . . . . . . . . . . . . . . . . . . . . . . 107
5.2 Sameness of structure and ante rem structuralism . . . . . . . . . 108
5.3 Interpretability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.4 Biinterpretability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.5 From structures to theories . . . . . . . . . . . . . . . . . . . . . . 114
5.6 Interpretability and the transfer of truth . . . . . . . . . . . . . . . 119
5.7 Interpretability and arithmetical equivalence . . . . . . . . . . . . 123
5.8 Interpretability and transfer of proof . . . . . . . . . . . . . . . . . 126
5.9 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
5.a Arithmetisation of syntax and incompleteness . . . . . . . . . . . 130
5.b Definitional equivalence in second-order logic . . . . . . . . . . . 132

B Categoricity 137

6 Modelism and mathematical doxology 143


6.1 Towards modelism . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
contents xiii

6.2 Objects-modelism . . . . . . . . . . . . . . . . . . . . . . . . . . . 144


6.3 Doxology, objectual version . . . . . . . . . . . . . . . . . . . . . . 145
6.4 Concepts-modelism . . . . . . . . . . . . . . . . . . . . . . . . . . 146
6.5 Doxology, conceptual version . . . . . . . . . . . . . . . . . . . . . 148

7 Categoricity and the natural numbers 151


7.1 Moderate modelism . . . . . . . . . . . . . . . . . . . . . . . . . . 151
7.2 Aspirations to Categoricity . . . . . . . . . . . . . . . . . . . . . . 153
7.3 Categoricity within first-order model theory . . . . . . . . . . . . 153
7.4 Dedekind’s Categoricity Theorem . . . . . . . . . . . . . . . . . . 154
7.5 Metatheory of full second-order logic . . . . . . . . . . . . . . . . 155
7.6 Attitudes towards full second-order logic . . . . . . . . . . . . . . 156
7.7 Moderate modelism and full second-order logic . . . . . . . . . . 158
7.8 Clarifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
7.9 Moderation and compactness . . . . . . . . . . . . . . . . . . . . . 161
7.10 Weaker logics which deliver categoricity . . . . . . . . . . . . . . . 162
7.11 Application to specific kinds of moderate modelism . . . . . . . . 164
7.12 Two simple problems for modelists . . . . . . . . . . . . . . . . . 167
7.a Proof of the Löwenheim–Skolem Theorem . . . . . . . . . . . . . 167

8 Categoricity and the sets 171


8.1 Transitive models and inaccessibles . . . . . . . . . . . . . . . . . 171
8.2 Models of first-order set theory . . . . . . . . . . . . . . . . . . . . 173
8.3 Zermelo’s Quasi-Categoricity Theorem . . . . . . . . . . . . . . . 178
8.4 Attitudes towards full second-order logic: redux . . . . . . . . . . 179
8.5 Axiomatising the iterative process . . . . . . . . . . . . . . . . . . 182
8.6 Isaacson and incomplete structure . . . . . . . . . . . . . . . . . . 184
8.a Zermelo Quasi-Categoricity . . . . . . . . . . . . . . . . . . . . . 186
8.b Elementary Scott–Potter foundations . . . . . . . . . . . . . . . . 192
8.c Scott–Potter Quasi-Categoricity . . . . . . . . . . . . . . . . . . . 197

9 Transcendental arguments against model-theoretical scepticism 203


9.1 Model-theoretical scepticism . . . . . . . . . . . . . . . . . . . . . 203
9.2 Moorean versus transcendental arguments . . . . . . . . . . . . . 206
9.3 The Metaresources Transcendental Argument . . . . . . . . . . . 206
9.4 The Disquotational Transcendental Argument . . . . . . . . . . . 210
9.5 Ineffable sceptical concerns . . . . . . . . . . . . . . . . . . . . . . 214
9.a Application: the (non-)absoluteness of truth . . . . . . . . . . . . 217

10 Internal categoricity and the natural numbers 223


10.1 Metamathematics without semantics . . . . . . . . . . . . . . . . 224
xiv contents

10.2 The internal categoricity of arithmetic . . . . . . . . . . . . . . . . 227


10.3 Limits on what internal categoricity could show . . . . . . . . . . 229
10.4 The intolerance of arithmetic . . . . . . . . . . . . . . . . . . . . . 232
10.5 A canonical theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
10.6 The algebraic / univocal distinction . . . . . . . . . . . . . . . . . 233
10.7 Situating internalism in the landscape . . . . . . . . . . . . . . . . 236
10.8 Moderate internalists . . . . . . . . . . . . . . . . . . . . . . . . . 237
10.a Connection to Parsons . . . . . . . . . . . . . . . . . . . . . . . . 239
10.b Proofs of internal categoricity and intolerance . . . . . . . . . . . 242
10.c Predicative Comprehension . . . . . . . . . . . . . . . . . . . . . . 246

11 Internal categoricity and the sets 251


11.1 Internalising Scott–Potter set theory . . . . . . . . . . . . . . . . . 251
11.2 Quasi-intolerance for pure set theory . . . . . . . . . . . . . . . . 253
11.3 The status of the continuum hypothesis . . . . . . . . . . . . . . . 255
11.4 Total internal categoricity for pure set theory . . . . . . . . . . . . 256
11.5 Total intolerance for pure set theory . . . . . . . . . . . . . . . . . 257
11.6 Internalism and indefinite extensibility . . . . . . . . . . . . . . . 258
11.a Connection to McGee . . . . . . . . . . . . . . . . . . . . . . . . . 260
11.b Connection to Martin . . . . . . . . . . . . . . . . . . . . . . . . . 262
11.c Internal quasi-categoricity for SP . . . . . . . . . . . . . . . . . . . 263
11.d Total internal categoricity for CSP . . . . . . . . . . . . . . . . . . 266
11.e Internal quasi-categoricity of ordinals . . . . . . . . . . . . . . . . 268

12 Internal categoricity and truth 271


12.1 The promise of truth-internalism . . . . . . . . . . . . . . . . . . . 271
12.2 Truth operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
12.3 Internalism about model theory and internal realism . . . . . . . . 276
12.4 Truth in higher-order logic . . . . . . . . . . . . . . . . . . . . . . 282
12.5 Two general issues for truth-internalism . . . . . . . . . . . . . . . 284
12.a Satisfaction in higher-order logic . . . . . . . . . . . . . . . . . . . 285

13 Boolean-valued structures 295


13.1 Semantic-underdetermination via Push-Through . . . . . . . . . . 295
13.2 The theory of Boolean algebras . . . . . . . . . . . . . . . . . . . . 296
13.3 Boolean-valued models . . . . . . . . . . . . . . . . . . . . . . . . 298
13.4 Semantic-underdetermination via filters . . . . . . . . . . . . . . . 301
13.5 Semanticism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
13.6 Bilateralism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
13.7 Open-ended-inferentialism . . . . . . . . . . . . . . . . . . . . . . 311
13.8 Internal-inferentialism . . . . . . . . . . . . . . . . . . . . . . . . . 314
contents xv

13.9 Suszko’s Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316


13.a Boolean-valued structures with filters . . . . . . . . . . . . . . . . 321
13.b Full second-order Boolean-valued structures . . . . . . . . . . . . 323
13.c Ultrafilters, ultraproducts, Łoś, and compactness . . . . . . . . . . 326
13.d The Boolean-non-categoricity of CBA . . . . . . . . . . . . . . . . 328
13.e Proofs concerning bilateralism . . . . . . . . . . . . . . . . . . . . 330

C Indiscernibility and classification 333

14 Types and Stone spaces 337


14.1 Types for theories . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
14.2 An algebraic view on compactness . . . . . . . . . . . . . . . . . . 338
14.3 Stone’s Duality Theorem . . . . . . . . . . . . . . . . . . . . . . . 339
14.4 Types, compactness, and stability . . . . . . . . . . . . . . . . . . 342
14.5 Bivalence and compactness . . . . . . . . . . . . . . . . . . . . . . 346
14.6 A biinterpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
14.7 Propositions and possible worlds . . . . . . . . . . . . . . . . . . . 350
14.a Topological background . . . . . . . . . . . . . . . . . . . . . . . . 354
14.b Bivalent-calculi and bivalent-universes . . . . . . . . . . . . . . . . 356

15 Indiscernibility 359
15.1 Notions of indiscernibility . . . . . . . . . . . . . . . . . . . . . . 359
15.2 Singling out indiscernibles . . . . . . . . . . . . . . . . . . . . . . 366
15.3 The identity of indiscernibles . . . . . . . . . . . . . . . . . . . . . 370
15.4 Two-indiscernibles in infinitary logics . . . . . . . . . . . . . . . . 376
15.5 n-indiscernibles, order, and stability . . . . . . . . . . . . . . . . . 380
15.a Charting the grades of discernibility . . . . . . . . . . . . . . . . . 384

16 Quantifiers 387
16.1 Generalised quantifiers . . . . . . . . . . . . . . . . . . . . . . . . 387
16.2 Clarifying the question of logicality . . . . . . . . . . . . . . . . . 389
16.3 Tarski and Sher . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
16.4 Tarski and Klein’s Erlangen Programme . . . . . . . . . . . . . . . 390
16.5 The Principle of Non-Discrimination . . . . . . . . . . . . . . . . 392
16.6 The Principle of Closure . . . . . . . . . . . . . . . . . . . . . . . 399
16.7 McGee’s squeezing argument . . . . . . . . . . . . . . . . . . . . . 407
16.8 Mathematical content . . . . . . . . . . . . . . . . . . . . . . . . . 408
16.9 Explications and pluralism . . . . . . . . . . . . . . . . . . . . . . 410

17 Classification and uncountable categoricity 413


xvi contents

17.1 The nature of classification . . . . . . . . . . . . . . . . . . . . . . 413


17.2 Shelah on classification . . . . . . . . . . . . . . . . . . . . . . . . 419
17.3 Uncountable categoricity . . . . . . . . . . . . . . . . . . . . . . . 426
17.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432
17.a Proof of Proposition 17.2 . . . . . . . . . . . . . . . . . . . . . . . . 432

D Historical appendix 435

18 Wilfrid Hodges
A short history of model theory 439
18.1 ‘A new branch of metamathematics’ . . . . . . . . . . . . . . . . . 439
18.2 Replacing the old metamathematics . . . . . . . . . . . . . . . . . 440
18.3 Definable relations in one structure . . . . . . . . . . . . . . . . . 445
18.4 Building a structure . . . . . . . . . . . . . . . . . . . . . . . . . . 449
18.5 Maps between structures . . . . . . . . . . . . . . . . . . . . . . . 455
18.6 Equivalence and preservation . . . . . . . . . . . . . . . . . . . . . 460
18.7 Categoricity and classification theory . . . . . . . . . . . . . . . . 465
18.8 Geometric model theory . . . . . . . . . . . . . . . . . . . . . . . 469
18.9 Other languages . . . . . . . . . . . . . . . . . . . . . . . . . . . . 472
18.10 Model theory within mathematics . . . . . . . . . . . . . . . . . . 474
18.11 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475
18.12 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . 475

Bibliography 477

Index 507

Index of names 513

Index of symbols and definitions 515


A
Reference and realism
Introduction to Part A

The two central themes of Part A are reference and realism.


Here is an old philosophical chestnut: How do we (even manage to) represent the
world? Our most sophisticated representations of the world are perhaps linguistic.
So a specialised—but still enormously broad—version of this question is: How do
words (even manage to) represent things?
Enter model theory. One of the most basic ideas in model theory is that a struc-
ture assigns interpretations to bits of vocabulary, and in such a way that we can make
excellent sense of the idea that the structure makes each sentence (in that vocabu-
lary) either true or false. Squint slightly, and model theory seems to be providing us
with a perfectly precise, formal way to understand certain aspects of linguistic rep-
resentation. It is no surprise at all, then, that almost any philosophical discussion
of linguistic representation, or reference, or truth, ends up invoking notions which
are recognisably model-theoretic.
In Chapter 1, we introduce the building blocks of model theory: the notions of
signature, structure, and satisfaction. Whilst the bare technical bones should be
familiar to anyone who has covered a 101-level course in mathematical logic, we
also discuss the philosophical question: How should we best understand quantifiers
and variables? Here we see that philosophical issues arise at the very outset of our
model-theoretic investigations. We also introduce second-order logic and its var-
ious semantics. While second-order logic is less commonly employed in contem-
porary model theory, it is employed frequently in philosophy of model theory, and
understanding the differences between its various semantics will be important in
many subsequent chapters.
In Chapter 2, we examine various concerns about the determinacy of reference
and so, perhaps, the determinacy of our representations. Here we encounter fa-
mous arguments from Benacerraf and Putnam, which we explicate using the for-
mal Push-Through Construction. Since isomorphic structures are elementarily
equivalent—that is, they make exactly the same sentences true and false—this
threatens the conclusion that it is radically indeterminate, which of many isomor-
phic structures accurately captures how language represents the world.
Now, one might think that the reference of our word ‘cat’ is constrained by the
causal links between cats and our uses of that word. Fair enough. But there are no
causal links between mathematical objects and mathematical words. So, on certain
conceptions of what humans are like, we will be unable to answer the question:
How do we (even manage to) refer to any particular mathematical entity? That is, we
will have to accept that we do not refer to particular mathematical entities.
4 introduction to part a

Whilst discussing these issues, we introduce Putnam’s famous just-more-theory


manoeuvre. It is important to do this both clearly and early, since many versions of
this dialectical move occur in the philosophical literature on model theory. Indeed,
they occur especially frequently in Part B of this book.
Now, philosophers have often linked the topic of reference to the topic of real-
ism. One way to draw the connection is as follows: If reference is radically inde-
terminate, then my word ‘cabbage’ and my word ‘cat’ fail to pick out anything de-
terminately. So when I say something like ‘there is a cabbage and there is a cat’, I
have at best managed to say that there are at least two distinct objects. That seems
to fall far short of expressing any real commitment to cats and cabbages themselves.1
In short, radical referential indeterminacy threatens to undercut certain kinds of
realism altogether. But only certain kinds: we close Chapter 2 by suggesting that
some versions of mathematical platonism can live with the fact that mathematical
language is radically referentially indeterminate by embracing a supervaluational
semantics.
Concerns about referential indeterminacy also feature in discussions about real-
ism within the philosophy of science. In Chapter 3, we examine a particular version
of scientific realism that arises by considering Ramsey sentences. Roughly, these
are sentences where all the ‘theoretical vocabulary’ has been ‘existentially quanti-
fied away’. Ramsey sentences seem promising, since they seem to incur a kind of
existential commitment to theoretical entities, which is characteristic of realism,
whilst making room for a certain level referential indeterminacy. We look at the re-
lation between Newman’s objection and the Push-Through Construction of Chap-
ter 2, and between Ramsey sentences and various model-theoretic notions of con-
servation. Ultimately, by combining the Push-Through Construction with these
notions of conservation, we argue that the dialectic surrounding Newman’s objec-
tion should track the dialectic of Chapter 2, surrounding Putnam’s permutation ar-
gument in the philosophy of mathematics.
The notions of conservation we introduce in Chapter 3 are crucial to Abraham
Robinson’s attempt to use model theory to salvage Leibniz’s notion of an ‘infinites-
imal’. Infinitesimals are quantities whose absolute value is smaller than that of any
given positive real number. They were an important part of the historical calculus;
they fell from grace with the rise of ε–δ notation; but they were given a new lease
of life within model theory via Robinson’s non-standard analysis. This is the topic
of Chapter 4. Here we introduce the idea of compactness to prove that the use of
infinitesimals is consistent.
Robinson believed that this vindicated the viability of the Leibnizian approach
to the calculus. Against this, Bos has questioned whether Robinson’s non-standard
analysis is genuinely faithful to Leibniz’s mathematical practice. In Chapter 4, we

1 Cf. Putnam (1977: 491) and Button (2013: 59–60).


introduction to part a 5

offer a novel defence of Robinson. By building valuations into Robinson’s model


theory, we prove new results which allow us to approximate more closely what we
know about the Leibnizian conception of the structure of the infinitesimals. In-
deed, we show that Robinson’s non-standard analysis can rehabilitate various his-
torical methods for reasoning with and about infinitesimals that have fallen far from
fashion.
The question remains, of whether we should believe in infinitesimals. Leib-
niz himself was tempted to treat his infinitesimals as ‘convenient fictions’; Robin-
son explicitly regarded his infinitesimals in the same way; and their method of in-
troduction in model theory allows for perhaps the cleanest possible version of a
fictionalist-cum-instrumentalist attitude towards ‘troublesome’ entities. Indeed,
we can prove that reasoning as if there are infinitesimals will only generate results
that one could have obtained without that assumption. One can have anti-realism,
then, with a clear conscience.
In Chapter 5, we take a step back from these specific applications of model the-
ory, to discuss a more methodological question about the philosophical application
of model theory: under what circumstances should we call two structures ‘the same’?
This question can be posed within mathematics, where its answer will depend upon
the similarities and differences that matter for the mathematical purposes at hand.
But the question can also be given a metaphysical gloss. In particular, consider a
philosopher who thinks (for example) that: (a) there is a single, abstract, entity
which is ‘the natural number structure’, and that (b) there is a single, abstract en-
tity which is ‘the structure of the integers’; but that (c) these two entities are dis-
tinct. Then this philosopher must provide an account of identity and distinctness
between ‘structures’, so construed; and we show just how hard this is.
Notions of sameness of structure also induce notions of sameness of theory. Af-
ter surveying a wide variety of formal notions of sameness of structure and theory,
we discuss three ambitious claims concerning what sameness of theory preserves,
namely: truth; arithmetical provability; and proof. We conclude that more philo-
sophically ambitious versions of these preservation-theses generally fail.
This meta-issue of sameness of structure and theory is a good place to end Part
A, though, both because (a) the discussion is enhanced by the specific examples
of structures and theories discussed earlier in the text, and because (b) questions
about sameness of structure and theory inform a number of the discussions and
debates which we treat in later Parts of the book.
Readers who only want to dip into particular topics of Part A can consult the
following Hasse diagram of dependencies between the sections of Part A, whilst
referring to the table of contents. A section y depends upon a section x iff there is
a path leading downwards from x to y. So, a reader who wants to get straight to the
discussion of fictionalism about infinitesimals will want to leap straight to §4.7; but
6 introduction to part a

they should know that this section assumes a prior understanding of §§2.1, 4.1, 4.2,
and much (but not all) of Chapter 1. (We omit purely technical appendices from
this diagram.)

1.1
1.2
1.3
1.5 1.4
1.7 1.6
1.9 1.8
1.10
2.1 1.11
2.2 1.12 1.a
2.3 3.1 1.13 1.b 1.c
2.4 2.a 3.2 4.1
2.5 3.3 4.2
3.5 4.3
3.6 4.4
5.1 3.4 3.7 4.5 4.7
5.2 5.3 3.8 4.6
5.4 3.9 4.8
5.5 3.10 3.a 4.9
5.6 5.7 5.8 4.10
1
Logics and languages

Model theory begins by considering the relationship between languages and struc-
tures. This chapter outlines the most basic aspects of that relationship.
One purpose of the chapter will therefore be immediately clear: we want to lay
down some fairly dry, technical preliminaries. Readers with some familiarity with
mathematical logic should feel free to skim through these technicalities, as there are
no great surprises in store.
Before the skimming commences, though, we should flag a second purpose of
this chapter. There are at least three rather different approaches to the semantics
for formal languages. In a straightforward sense, these approaches are technically
equivalent. Most books simply choose one of them without comment. We, how-
ever, lay down all three approaches and discuss their comparative strengths and
weaknesses. Doing this highlights that there are philosophical discussions to be had
from the get-go. Moreover, by considering what is invariant between the different
approaches, we can better distinguish between the merely idiosyncratic features of
a particular approach, and the things which really matter.
One last point, before we get going: tradition demands that we issue a caveat.
Since Tarski and Quine, philosophers have been careful to emphasise the impor-
tant distinction between using and mentioning words. In philosophical texts, that
distinction is typically flagged with various kinds of quotation marks. But within
model theory, context almost always disambiguates between use and mention.
Moreover, including too much punctuation makes for ugly text. With this in mind,
we follow model-theoretic practice and avoid using quotation marks except when
they will be especially helpful.

1.1 Signatures and structures


We start with the idea that formal languages can have primitive vocabularies:

Definition 1.1: A signature, L , is a set of symbols, of three basic kinds: constant sym-
bols, relation symbols, and function symbols. Each relation symbol and function symbol
has an associated number of places (a natural number), so that one may speak of an
n-place relation or function symbol.
8 logics and languages

Throughout this book, we use script fonts for signatures. Constant symbols should
be thought of as names for entities, and we tend to use c 1 , c 2 , etc. Relation symbols,
which are also known as predicates, should be thought of as picking out properties
or relations. A two-place relation, such as x is smaller than y, must be associated
with a two-place relation symbol. We tend to use R1 , R2 , etc. for relation symbols.
Function symbols should be thought of as picking out functions and, again, they
need an associated number of places: the function of multiplication on the natural
numbers takes two natural numbers as inputs and outputs a single natural number,
so we must associate that function with a two-place function symbol. We tend to
use f 1 , f 2 , etc. for function symbols.
The examples just given—being smaller than, and multiplication on the natural
numbers—suggest that we will use our formal vocabulary to make determinate
claims about certain objects, such as people or numbers. To make this precise,
we introduce the notion of an L -structure; that is, a structure whose signature is
L . An L -structure, M, is an underlying domain, M, together with an assignment
of L ’s constant symbols to elements of M, of L ’s relation symbols to relations
on M, and of L ’s function symbols to functions over M. We always use calligraphic
fonts M, N, … for structures, and M, N, … for their underlying domains. Where s
is any L -symbol, we say that s M is the object, relation or function (as appropriate)
assigned to s in the structure M. This informal explanation of an L -structure is
always given a set-theoretic implementation, leading to the following definition:

Definition 1.2: An L -structure, M, consists of:


• a non-empty set, M, which is the underlying domain of M,
• an object c M ∈ M for each constant symbol c from L ,
• a relation RM ⊆ M n for each n-place relation symbol R from L , and
• a function f M : M n Ð→ M for each n-place function symbol f from L .

As is usual in set theory, M n is just the set of n-tuples over M, i.e.:1

M n = {(a1 , …, an ) : a1 ∈ M and … and an ∈ M}

Likewise, we implement a function g : M n Ð→ M in terms of its set-


theoretic graph. That is, g will be a subset of M n+1 such that if (x1 , …, xn , y) and
(x1 , …, xn , z) are elements of g then y = z and such that for every (x1 , …, xn ) in M n
there is y in M such that (x1 , …, xn , y) is in g. But we continue to think about func-
tions in the normal way, as maps sending n-tuples of the domain, M n , to elements
of the co-domain, M, so tend to write (x1 , …, xn , y) ∈ g just as g(x1 , …, xn ) = y.
1 The full definition of X n is by recursion: X 1 = X and X n+1 = X n × X, where A × B = {(a, b) :

a ∈ A and b ∈ B}. Likewise, we recursively define ordered n-tuples in terms of ordered pairs by setting e.g.
(a, b, c) = ((a, b), c).
1.2. first-order logic: a first look 9

Given the set-theoretic background, L -structures are individuated extension-


ally: they are identical iff they have exactly the same underlying domain and make
exactly the same assignments. So, where M, N are L -structures, M = N iff both
M = N and s M = s N for all s from L . To obtain different structures, then, we can
either change the domain, change the interpretation of some symbol(s), or both.
Structures are, then, individuated rather finely, and indeed we will see in Chapters
2 and 5 that this individuation is too fine for many purposes. But for now, we can
simply observe that there are many, many different structures, in the sense of Defi-
nition 1.2.

1.2 First-order logic: a first look


We know what (L -)structures are. To move to the idea of a model, we need to think
of a structure as making certain sentences true or false. So we must build up to the
notion of a sentence. We start with their syntax.

Syntax for first-order logic


Initially, we restrict our attention to first-order sentences. These are the sentences we
obtain by adding a basic starter-pack of logical symbols to a signature (in the sense
of Definition 1.1). These logical symbols are:
• variables: u, v, w, x, y, z, with numerical subscripts as necessary
• the identity sign: =
• a one-place sentential connective: ¬
• two-place sentential connectives: ∧, ∨
• quantifiers: ∃, ∀
• brackets: (, )
We now offer a recursive definition of the syntax of our language:2

Definition 1.3: The following, and nothing else, are first-order L -terms:
• any variable, and any constant symbol c from L

2 A pedantic comment is in order. The symbols ‘t ’ and ‘t ’ are not being used here as expressions in
1 2
the object language (i.e. first-order logic with signature L ). Rather, they are being used as expressions of
the metalanguage, within which we describe the syntax of first-order L -terms and L -formulas. Similarly,
the symbol ‘x’, as it occurs in the last clause of Definition 1.3, is not being used as an expression of the object
language, but in the metalanguage. So the final clause in this definition should be read as saying something
like this. For any variable and any formula φ which does not already contain a concatenation of a quantifier
followed by that variable, the following concatenation is a formula: a quantifier, followed by that variable,
followed by φ. (The reason for this clause is to guarantee that e.g. ∃v∀vF(v) is not a formula.) We could
flag this more explicitly, by using a different font for metalinguistic variables (for example). However, as
with flagging quotation, we think the additional precision is not worth the ugliness.
10 logics and languages

• f (t 1 , …, t n ), for any L -terms t 1 , …, t n and any n-place function symbol f


from L
The following, and nothing else, are first-order L -formulas:
• t 1 = t 2 , for any L -terms t 1 and t 2
• R(t 1 , …, t n ), for any L -terms t 1 , …, t n and any n-place relation symbol R
from L
• ¬φ, for any L -formula φ
• (φ ∧ ψ) and (φ ∨ ψ), for any L -formulas φ and ψ
• ∃xφ and ∀xφ, for any variable x and any L -formula φ which contains neither
of the expressions ∃x nor ∀x.
Formulas of the first two sorts—i.e. terms appropriately concatenated either with the
identity sign or an L -predicate—are called atomic L -formulas.

As is usual, for convenience we add two more sentential connectives, → and ↔,


with their usual abbreviations. So, (φ → ψ) abbreviates (¬φ ∨ ψ), and (φ ↔ ψ)
abbreviates ((φ → ψ) ∧ (ψ → φ)). We will also use some extremely common
bracketing conventions to aid readability, so we sometimes use square brackets
rather than rounded brackets, and we sometimes omit brackets where no ambiguity
can arise.
We say that a variable is bound if it occurs within the scope of a quantifier, i.e. we
have something like ∃x(…x…). A variable is free if it is not bound. We now say
that an L -sentence is an L -formula containing no free variables. When we want
to draw attention to the fact that some formula φ has certain free variables, say x
and y, we tend to do this by writing the formula as φ(x, y). We say that φ(x, y) is a
formula with free variables displayed iff x and y are the only free variables in φ. When
we consider a sequence of n-variables, such as v1 , …, vn , we usually use overlining
to write this more compactly, as v, leaving it to context to determine the number
of variables in the sequence. So if we say ‘φ(x) is a formula with free variables dis-
played’, we mean that all and only its free variables are in the sequence x. We also use
similar overlining for other expressions. For example, we could have phrased part
of Definition 1.3 as follows: f (t) is a term whenever each entry in t is an L -term
and f is a function symbol from L .

Semantics: the trouble with quantifiers


We now understand the syntax of first-order sentences. Later, we will consider log-
ics with a more permissive syntax. But first-order logic is something like the default,
for both philosophers and model theorists. And our next task is to understand its
semantics. Roughly, our aim is to define a relation, ⊧, which obtains between a struc-
ture and a sentence just in case (intuitively) the sentence is true in the structure. In
1.2. first-order logic: a first look 11

fact, there are many different but extensionally equivalent approaches to defining
this relation, and we will consider three in this chapter.
To understand why there are several different approaches to the semantics for
first-order logic, we must see why the most obvious approach fails. Our sentences
have a nice, recursive syntax, so we will want to provide them with a nice, recursive
semantics. The most obvious starting point is to supply semantic clauses for the
two kinds of atomic sentence, as follows:

M ⊧ t 1 = t 2 iff t M
1 = t2
M

M ⊧ R(t 1 , …, t n ) iff (t M
1 , …, t n ) ∈ R
M M

Next, we would need recursion clauses for the quantifier-free sentences. So, writing
M ⊭ φ for it is not the case that M ⊧ φ, we would offer:

M ⊧ ¬φ iff M ⊭ φ
M ⊧ (φ ∧ ψ) iff M ⊧ φ and M ⊧ ψ

So far, so good. But the problem arises with the quantifiers. Where the notation
φ(c/x) indicates the formula obtained by replacing every instance of the free vari-
able x in φ(x) with the constant symbol c, an obvious thought would be to try:

M ⊧ ∀xφ(x) iff M ⊧ φ(c/x) for every constant symbol c from L

Unfortunately, this recursion clause is inadequate. To see why, suppose we had a very
simple signature containing a single one-place predicate R and no constant symbols.
Then, for any structure M in that signature, we would vacuously have that M ⊧
∀vR(v). But this would be the case even if RM = ∅, that is, even if nothing had the
property picked out by R. Intuitively, that is the wrong verdict.
The essential difficulty in defining the semantics for first-order logic therefore
arises when we confront quantifiers. The three approaches to semantics which we
consider present three ways to overcome this difficulty.

Why it is worth considering different approaches


In a straightforward sense, the three approaches are technically equivalent. So most
books simply adopt one of these approaches, without comment, and get on with
other things. In deciding to present all three approaches here, we seem to be trebling
our reader’s workload. So we should pause to explain our decision.
First: the three approaches to semantics are so intimately related, at a technical
level, that the workload is probably only doubled, rather than trebled.
Second: readers who are happy ploughing through technical definitions will find
nothing very tricky here. And such readers should find that the additional technical
12 logics and languages

investment gives a decent philosophical pay-off. For, as we move through the chap-
ter, we will see that these (quite dry) technicalities can both generate and resolve
philosophical controversies.
Third: we expect that even novice philosophers reading this book will have at
least a rough and ready idea of what is coming next. And such readers will be bet-
ter served by reading (and perhaps only partially absorbing) multiple different ap-
proaches to the semantics for first-order logic, than by trying to rote-learn one spe-
cific definition. They will thereby get a sense of what is important to supplying a
semantics, and what is merely an idiosyncratic feature of a particular approach.

1.3 The Tarskian approach to semantics


We begin with the Tarskian approach.3 Recall that the ‘obvious’ semantic clauses
fail because L may not contain enough constant symbols. The Tarskian approach
handles this problem by assigning interpretations to the variables of the language.
In particular, where M is any L -structure, a variable-assignment is any function σ
from the set of variables to the underlying domain M. We then define satisfaction
with respect to pairs of structures with variable-assignments.
To do this, we must first specify how the structure / variable-assignment pair de-
termines the behaviour of the L -terms. We do this by recursively defining an ele-
ment t M,σ of M for a term t with free variables among x1 , …, xn as follows:

t M,σ = σ(xi ), if t is the variable xi


t M,σ = f M (s M,σ
1 , …, s k ), if t is the term f (s 1 , …, s k )
M,σ

To illustrate this definition, suppose that M is the natural numbers in the signature
{0, 1, +, ×}, with each symbol interpreted as normal. (This licenses us in dropping
the ‘M’-superscript when writing the symbols.) Suppose that σ and τ are variable-
assignments such that σ(x1 ) = 5, σ(x2 ) = 7, τ(x1 ) = 3, τ(x2 ) = 7, and consider
the term t(x1 , x2 ) = (1 + x1 ) × (x1 + x2 ). Then we can compute the interpretation
of the term relative to the variable-assignments as follows:

t M,σ = (1 + xM,σ
1 ) × (x1
M,σ
+ xM,σ
2 ) = (1 + 5) × (5 + 7) = 72

t M,τ = (1 + xM,τ
1 ) × (x1
M,τ
+ xM,τ
2 ) = (1 + 3) × (3 + 7) = 40

We next define the notion of satisfaction relative to a variable-assignment:

3 See Tarski (1933) and Tarski and Vaught (1958), but also §12.a.
1.4. semantics for variables 13

M, σ ⊧ t 1 = t 2 iff t M,σ
1 = t M,σ
2 , for any L -terms t 1 , t 2

M, σ ⊧ R(t 1 , …, t n ) iff (t M,σ


1 , …, t n ) ∈ R , for any L -terms t 1 , …, t n
M,σ M

and any n-place relation symbol R from L


M, σ ⊧ ¬φ iff M, σ ⊭ φ
M, σ ⊧ (φ ∧ ψ) iff M, σ ⊧ φ and M, σ ⊧ ψ
M, σ ⊧ ∀xφ(x) iff M, τ ⊧ φ(x) for every variable-assignment τ
which agrees with σ except perhaps on the value of x

We leave it to the reader to formulate clauses for disjunction and existential quan-
tification. Finally, where φ is any first-order L -sentence, we say that M ⊧ φ iff
M, σ ⊧ φ for all variable-assignments σ.

1.4 Semantics for variables


The Tarskian approach is technically flawless. However, the apparatus of variable-
assignments raises certain philosophical issues.
A variable-assignment effectively gives variables a particular interpretation. In
that sense, variables are treated rather like names (or constant symbols). However,
when we encounter the clause for a quantifier binding a variable, we allow ourselves
to consider all of the other ways that the bound variable might have been interpreted.
In short, the Tarskian approach treats variables as something like varying names.
This gives rise to a philosophical question: should we regard variables as vary-
ing names? With Quine, our answer is No: ‘the “variation” connoted [by the word
“variable”] belongs to a vague metaphor which is best forgotten.’4
To explain why we say this, we begin with a simple observation. A Tarskian
variable-assignment may assign different semantic values to the formulas x > 0 and
y > 0. But, on the face of it, that seems mistaken. As Fine puts the point, using one
variable rather than the other ‘would appear to be as clear a case as any of a mere
“conventional” or “notational” difference; the difference is merely in the choice of
the symbol and not in its linguistic function.’5 And this leads Fine to say:
(a) ‘Any two variables (ranging over a given domain of objects) have the same
semantic role.’

4 Quine (1981: §12). For ease of reference, we cite the 1981-edition. However, the relevant sections

are entirely unchanged from the (first) 1940-edition. We owe several people thanks for discussion of ma-
terial in this section. Michael Potter alerted us to Bourbaki’s notation; Kai Wehmeier alerted us to Quine’s
(cf. Wehmeier forthcoming); and Robert Trueman suggested that we should connect all of this to Fine’s
antinomy of the variable.
5 Fine (2003: 606, 2007: 7), for this and all subsequent quotes from Fine.
14 logics and languages

But, as Fine notes, this cannot be right either. For, ‘when we consider the semantic
role of the variables in the same expression—such as “x > y”—then it seems equally
clear that their semantic role is different.’ So Fine says:
(b) ‘Any two variables (ranging over a given domain of objects) have a different
semantic role.’
And now we have arrived at Fine’s antinomy of the variable.
We think that this whole antinomy gets going from the mistaken assumption that
we can assign a ‘semantic role’ to a variable in isolation from the quantifier which
binds it.6 As Quine put the point more than six decades before Fine: ‘The vari-
ables […] serve merely to indicate cross-references to various positions of quantifi-
cation.’7 Quine’s point is that ∃x∀yφ(x, y) and ∃y∀xφ(y, x) are indeed just typo-
graphical variants, but that both are importantly different from ∀x∃yφ(x, y). And
to illustrate this graphically, Quine notes that we could use a notation which aban-
dons typographically distinct variables altogether. For example, instead of writing:

∃x∀y((φ(x, y) ∧ ∃zφ(x, z)) → φ(y, x))

we might have written:8

∃∀((φ(●, ●) ∧ ∃φ(●, ●)) → φ(●, ●))

Bourbaki rigorously developed Quine’s brief notational suggestion.9 And the re-
sulting Quine–Bourbaki notation is evidently just as expressively powerful as our or-
dinary notation. However, if we adopt the Quine–Bourbaki notation, then we will
not even be able to ask whether typographically distinct variables like ‘x’ and ‘y’ have
different ‘semantic roles’, and Fine’s antinomy will dissolve away.10
6 Fine (2003: 610–14, 2007: 12–16) considers this thought, but does not consider the present point.
7 Quine (1981: 69–70). See also Curry (1933: 389–90), Quine (1981: iv, 5, 71), Dummett (1981: ch.1),
Kaplan (1986: 244), Lavine (2000: 5–6), and Potter (2000: 64).
8 Quine (1981: §12).
9 Bourbaki (1954: ch.1), apparently independently. The slight difference is that Bourbaki uses

Hilbert’s epsilon operator instead of quantifiers.


10 Pickel and Rabern (2017: 148–52) consider and criticise the Quine–Bourbaki approach to Fine’s anti-

nomy. Pickel and Rabern assume that the Quine–Bourbaki approach will be coupled with Frege’s idea that
one obtains the predicate ‘( ) ≤ ( )’ by taking a sentence like ‘7 ≤ 7’ and deleting the names. They then
insist that Frege must distinguish between the case when ‘( ) ≤ ( )’ is regarded as a one-place predicate,
and the case where it is regarded as a two-place predicate. And they then maintain: ‘if Frege were to intro-
duce marks capable of typographically distinguishing between these predicates, then that mark would need
its own semantic significance, which in this context means designation.’ We disagree with the last part of
this claim. Brackets are semantically significant, in that ¬(φ ∧ ψ) is importantly different from (¬φ ∧ ψ);
but brackets do not denote. Fregeans should simply insist that any ‘marks’ on predicate-positions have a
similarly non-denotational semantic significance. After all, their ultimate purpose is just to account for the
different ‘cross-referencing’ in ∀x∃yφ(x, y) and ∀x∃yφ(y, x).
1.5. the robinsonian approach to semantics 15

To be clear, no one is recommending that we should adopt the Quine–Bourbaki


notation in practice: it would be hard to read and a pain to typeset. To dissolve
the antimony of the variable, it is enough to know that we could in principle have
adopted this notation.
But there is a catch. Just as this notation leaves us unable to formulate Fine’s
antinomy of the variable, it leaves us unable to define the notion of a variable-
assignment. So, until we can provide a non-Tarskian approach to semantics, which
does not essentially rely upon variable-assignments, we have no guarantee that we
could have adopted the Quine–Bourbaki notation, even in principle. Now, we can
of course use the Tarskian approach to supply a semantics for Quine–Bourbaki sen-
tences derivatively.11 But if we were to do that that, we would lose the right to say
that we could, in principle, have done away with typographically distinct variables
altogether, for we would still be relying upon them in our semantic machinery.
In sum, we want an approach to semantics which (unlike Tarski’s) accords vari-
ables with no more apparent significance than is suggested by the Quine–Bourbaki
notation. Fortunately, such approaches are available.

1.5 The Robinsonian approach to semantics


To recall: difficulties concerning the semantics for quantifiers arise because L may
not contain names for every object in the domain. One solution to this problem is
obvious: just add new constants. This was essentially Robinson’s approach.12
To define how to add new symbols, it is easiest to define how to remove them.
Given a structure M, its L -reduct is the L -structure we obtain by ignoring the in-
terpretation of the symbols in M’s signature which are not in L . More precisely:13

Definition 1.4: Let L + and L be signatures with L + ⊇ L . Let M be an L + -


structure. Then M’s L -reduct, N, is the unique L -structure with domain M such that
s N = s M for all s from L . We also say that M is a signature-expansion of N, and that
N is a signature-reduct of M.

In Quinean terms, the difference between a model and its reduct is not ontological
but ideological.14 We do not add or remove any entities from the domain; we just
add or remove some (interpretations of) symbols.

11 Where φ is any Quine–Bourbaki sentence, let φ fo be the sentence of first-order logic which results

by: (a) inserting the variable vn after the nth quantifier in φ, counting quantifiers from left-to-right; (b)
replacing each blob connected to the nth -quantifier with the variable vn and (c) deleting all the connecting
wires. Then say M ⊧ φ iff M ⊧ φfo , with M ⊧ φfo defined via the Tarskian approach.
12 A. Robinson (1951: 19–21), with a tweak that one finds in, e.g., Sacks (1972: ch.4).
13 Cf. Hodges (1993: 9ff) and Marker (2002: 31).
14 Quine (1951: 14).
16 logics and languages

We can now define the idea of ‘adding new constants for every member of the
domain’. The following definition explains how to add, for each element a ∈ M, a
new constant symbol, c a , which is taken to name a:

Definition 1.5: Let L be any signature. For any set M, L (M) is the signature obtained
by adding to L a new constant symbol c a for each a ∈ M. For any L -structure M with
domain M, we say that M○ is the L (M)-structure whose L -reduct is M and such that

a = a for all a ∈ M.
cM

Since M○ is flooded with constants, it is very easy to set up its semantics. We start
by defining the interpretation of the L (M)-terms which contain no variables:
○ ○ ○ ○
1 , …, s k ), if t is the variable-free L (M)-term f (s 1 , …, s k )
t M = f M (s M M

For each atomic first-order L (M)-sentence, we then define:


○ ○
M○ ⊧ t 1 = t 2 iff t M
1 = t 2 , for any variable-free L (M)-terms t 1 , t 2
M

○ ○ ○
M○ ⊧ R(t 1 , …, t n ) iff (t M
1 , …, t n ) ∈ R
M M
, for
any variable-free L (M)-terms t 1 , …, t n and
any n-place relation symbol R from L (M)

And finally we offer:

M○ ⊧ ¬φ iff M○ ⊭ φ
M○ ⊧ (φ ∧ ψ) iff M○ ⊧ φ and M○ ⊧ ψ
M○ ⊧ ∀xφ(x) iff M○ ⊧ φ(c a /x) for every a ∈ M

We now have what we want, in terms of M○ . And, since M○ is uniquely determined


by M, we can now extract what we really wanted: definitions concerning M itself.
Where φ(v) is a first-order L -formula with free variables displayed, and a are from
M, we define a three-place relation which, intuitively, says that φ(v) is true of the
entities a according to M. Here is the definition:

M ⊧ φ(a) iff M○ ⊧ φ(c a /v)

The notation φ(c/v) indicates the L (M)-formula obtained by substituting the


kth constant in the sequence c for the kth variable in the sequence v. So we have
defined a three-place relation between an L -formula, entities a, and a structure M,
in terms of a two-place relation between a structure M○ and an L (M)-formula.
For readability, we will write φ(c) instead of φ(c/v), where no confusion arises.
As a limiting case, a sentence is a formula with no free variables. So for each L -
sentence φ, our definition states that M ⊧ φ iff M○ ⊧ φ. And, intuitively, we can
read this as saying that φ is true in M.
1.6. straining the notion of ‘language’ 17

To complete the Robinsonian semantics, we will define something similar for


terms. So, where a are entities from M and t(v) is an L -term with free variables
displayed, we define a function t M : M n Ð→ M, by:

t M (a) = (t(c a /v))M

This completes the Robinsonian approach. And the approach carries no taint of
the antinomy of the variable, since it clearly accords variables with no more seman-
tic significance than is suggested by the Quine–Bourbaki notation. Indeed, it is
easy to give a Robinsonian semantics directly for Quine–Bourbaki sentences, via:
M○ satisfies a Quine–Bourbaki sentence beginning with ‘∀’ iff for every a ∈ M the
model M○ satisfies the Quine–Bourbaki sentence which results from replacing all
blobs connected to the quantifier with ‘c a ’ and then deleting the quantifier and the
connecting wires.

1.6 Straining the notion of ‘language’


For all its virtues, the Robinsonian approach has some eyebrow-raising features of
its own. To define satisfaction for the sentences of the first-order L -sentences, we
have considered the sentences in some other formal languages, namely, those with
signature L (M) for any L -structure M. These languages can be enormous. Let M
be an infinite L -structure, whose domain M has size κ for some very big cardinal
κ.15 Then L (M) contains at least κ symbols. Can such a beast really count as a
language, in any intuitive sense?
Of course, there is no technical impediment to defining these enormous lan-
guages. So, if model theory is just regarded as a branch of pure mathematics, then
there is no real reason to worry about any of this. But we might, instead, want model
theory to be regarded as a branch of applied mathematics, whose (idealised) sub-
ject matter is the languages and theories that mathematicians actually use. And if
we regard model theory that way, then we will not want our technical notion of a
‘language’ to diverge too far from the kinds of things which we would ordinarily
count as languages.
There is a second issue with the Robinsonian approach. In Definition 1.5, we
introduced a new constant symbol, c a , for each a ∈ M. But we did not say what,
exactly, the constant symbol c a is. Robinson himself suggested that the constant
c a should just be the object a itself.16 In that case, every object in M○ would name
itself. But this is both philosophically strange and also technically awkward.
On the philosophical front: we might want to consider a structure, W, whose
domain is the set of all living wombats. In order to work out which sentences are
15 As is standard, we use κ to denote a cardinal; see the end of §1.b for a brief review of cardinals.
16 A. Robinson (1951: 21).
18 logics and languages

true in W using Robinson’s own proposal, we would have to treat each wombat as a
name for itself, and so imagine a language whose syntactic parts are live wombats.17
This stretches the ordinary notion of a language to breaking point.
There is also a technical hitch with Robinson’s own proposal. Suppose that c is
a constant symbol of L . Suppose that M is an L -structure where the symbol c
is itself an element of M’s underlying domain. Finally, suppose that M interprets c
as naming some element other than c itself, i.e. c M ≠ c. Now Robinson’s proposal

requires that c M = c. But since M○ is a signature expansion of M, we require that

c M = c M , which is a contradiction.
To fix this bug whilst retaining Robinson’s idea that c a = a, we would have to
tweak the definition of an L -structure to ensure that the envisaged situation can-
not arise.18 A better alternative—which also spares the wombats—is to abandon
Robinson’s suggestion that c a = a, and instead define the symbol c a so that it is
guaranteed not to be an element of M’s underlying domain.19 So this is our official
Robinsonian semantics (even if it was not exactly Robinson’s).

1.7 The Hybrid approach to semantics


Tarskian and Robinsonian semantics are technically equivalent, in the following
sense: they use the same notion of an L -structure, they use the same notion of
an L -sentence, and they end up defining exactly the same relation, ⊧, between
structures and sentences. But, as we have seen, neither approach is exactly ideal. So
we turn to a third approach: a hybrid approach.
In the Robinsonian semantics, we used M○ to define the expression M ⊧ φ(a).
Intuitively, this states that φ is true of a in M. If we start by defining this notation—
which we can do quite easily—then we can use it to present a semantics with the
following recursion clauses:

1 = t 2 , for any variable-free L -terms t 1 , t 2


M ⊧ t 1 = t 2 iff t M M

M ⊧ R(t 1 , …, t n ) iff (t M
1 , …, t n ) ∈ R , for any variable-free L -terms
M M

t 1 , …, t n and any n-place relation symbol R from L


M ⊧ ¬φ iff M ⊭ φ
M ⊧ (φ ∧ ψ) iff M ⊧ φ and M ⊧ ψ
M ⊧ ∀vφ(v) iff M ⊧ φ(a) for all a ∈ M

17 Cf. Lewis (1986: 145) on ‘Lagadonian languages’.


18 We would have to add a clause: if M is an L -structure and s ∈ L ∩ M, then s M = s.
19 A simple way to do this is as follows: let c be the ordered pair (a, M). By Foundation in the back-
a
ground set theory within which we implement our model theory, (a, M) ∉ M.
1.8. linguistic compositionality 19

All that remains is to define M ⊧ φ(a) without going all-out Robinsonian. And the
idea here is quite simple: we just add new constant symbols when we need them,
but not before. Here is the idea, rigorously developed. Let M be an L -structure
with a from M. For each ai among a, let c ai be a constant symbol not occurring in
L . Intuitively, we interpret each c ai as a name for ai . More formally, we define
M[a] to be a structure whose signature is L together with the new constant sym-
M[a]
bols among c a , whose L -reduct is M, and such that c ai = ai for each i. Where
φ(v) is an L -formula with free variables displayed, the Hybrid approach defines:

M ⊧ φ(a) iff M[a] ⊧ φ(c a /v)

When we combine our new definition of M ⊧ φ(a) with the clause for universal
quantification, we see that universal quantification effectively amounts to consider-
ing all the different ways of expanding the signature of M with a new constant sym-
bol which could be interpreted to name any element of M. (So the Hybrid approach
offers a semantics by simultaneous recursion over structures and languages.) Fi-
nally, we offer a similar clause for terms:

t M (a) = t M[a] (c a /v)

thereby completing the Hybrid approach.20

1.8 Linguistic compositionality


Unsurprisingly, the Hybrid approach is technically equivalent to the Robinsonian
and Tarskian approaches. However, its philosophical merits come out when we
revisit some of the potential defects of the other approaches. The Tarskian ap-
proach does not distinguish sufficiently between names and variables; the Hybrid
approach has no such issues. Indeed, just like the Robinsonian approach, the Hy-
brid approach accords variables with no greater semantic significance than is sug-
gested by the Quine–Bourbaki notation. But the Robinsonian approach involved
vast, peculiar ‘languages’; the Hybrid approach has no such issues. And, following
Lavine, we will pause on this last point.21
It is common to insist that languages should be compositional, in some sense. One
of the most famous arguments to this effect is due to Davidson. Because natural
languages are learnable, Davidson insists that ‘the meaning of each sentence [must
be] a function of a finite number of features of the sentence’. For, on the one hand,
20 The hybrid approach is hinted at by Geach (1962: 160), and Mates (1965: 54–7) offers something

similar. But the clearest examples we can find are Boolos and Jeffrey (1974: 104–5), Boolos (1975: 513–4),
and Lavine (2000: 10–12).
21 See Lavine’s (2000: 12–13) comments on compositionality and learnability.
20 logics and languages

if a language has this feature, then we ‘understand how an infinite aptitude can be
encompassed by finite accomplishments’. Conversely, if ‘a language lacks this fea-
ture then no matter how many sentences a would-be speaker learns to produce and
understand, there will remain others whose meanings are not given by the rules
already mastered.’22
Davidson’s argument is too quick. After all, it is a wild idealisation to suggest that
any actual human can indeed understand or learn the meanings of infinitely many
sentences: some sentences are just too long for any actual human to parse. It is
unclear, then, why we should worry about the ‘learnability’ of such sentences.
Still, something in the vicinity of Davidson’s argument seems right. In §1.6, we
floated the idea that model theory should be regarded as a branch of applied math-
ematics, whose (idealised) subject matter is the languages and theories that (pure)
mathematicians actually use. But here is an apparent phenomenon concerning that
subject matter: once we have a fixed interpretation in mind, we tend to act as if that
interpretation fixes the truth value of any sentence of the appropriate language, no
matter how long or complicated that sentence is.23 All three of our approaches to
formal semantics accommodate this point. For, given a signature L and an L -
structure M—i.e. an interpretation of the range of quantification and an interpre-
tation of each L -symbol—the semantic value of every L -sentence is completely
determined within M, in the sense that, for every L -sentence φ, either M ⊧ φ, or
M ⊧ ¬φ, but not both.
But the Hybrid approach, specifically, may allow us to go a little further. For,
when L is finite,24 and we offer the Hybrid approach to semantics, we may gain
some insight into how a finite mind might fully understand the rules by which an
interpretation fixes the truth-value of every sentence. That understanding seems to
reduce to three rather tractable components:
(a) an understanding of the finitely many recursion clauses governing satisfac-
tion for atomic sentences (finitely many, as we assumed that L is finite);
(b) an understanding of the handful of recursion clauses governing sentential
connectives; and
(c) an understanding of the recursion clauses governing quantification
On the Hybrid approach, point (c) reduces to an understanding of two ideas: (i)
the general idea that names can pick out objects,25 and (ii) the intuitive idea that, for
any object, we could expand our language with a new name for that object. In short:

22 Davidson (1965: 9).


23 A theme of Part B is whether, in certain circumstances, axioms can also fix truth values.
24 We can make a similar point if L can be recursively specified.
25 There are some deep philosophical issues concerning the question of how names pick out objects (see

Chapters 2 and 15). However, the general notion seems to be required by any model-theoretic semantics,
so that there is no special problem here for the Hybrid approach.
1.9. second-order logic: syntax 21

the Hybrid semantics seems to provide a truly compositional notion of meaning. But
we should be clear on what this means.
First, we are not aiming to escape what Sheffer once called the ‘logocentric
predicament’, that ‘In order to give an account of logic, we must presuppose and employ
logic.’26 Our semantic clause for object-language conjunction, ∧, always involves
conjunction in the metalanguage. On the Hybrid approach, our semantic clause
for object-language universal quantification, ∀, involved (metalinguistic) quantifi-
cation over all the ways in which a new name could be added to a signature. We do
not, of course, claim that anyone could read these semantic clauses and come to un-
derstand the very idea of conjunction or quantification from scratch. We are making
a much more mundane point: to understand the hybrid approach to semantics, one
need only understand a tractable number of ideas.
Second, in describing our semantics as compositional, we are not aiming to sup-
ply a semantics according to which the meaning of ∀xF(x) depends upon the sep-
arate meanings of the expressions ∀, x, F, and x.27 Not only would that involve an
oddly inflexible understanding of the word ‘compositional’; the discussion of §1.4
should have convinced us that variables do not have semantic values in isolation.28
Instead, on the hybrid approach, the meaning of ∀xF(x) depends upon the mean-
ings of the quantifier-expression ∀x…x and the predicate-expression F( ). The
crucial point is this: the Hybrid approach delivers the truth-conditions of infinitely
many sentences using only a small ‘starter pack’ of principles.
Having aired the virtues of the Hybrid approach, though, it is worth repeating
that our three semantic approaches are technically equivalent. As such, we can in
good faith use whichever approach we like, whilst claiming all of the pleasant philo-
sophical features of the Hybrid approach. Indeed, in the rest of this book, we simply
use whichever approach is easiest for the purpose at hand.
This concludes our discussion of first-order logic. It also concludes the ‘philo-
sophical’ component of this chapter. The remainder of this chapter sets down the
purely technical groundwork for several later philosophical discussions.

1.9 Second-order logic: syntax


Having covered first-order logic, we now consider second-order logic. This is much
less popular than first-order logic among working model-theorists. However, it has

26 Sheffer (1926: 228).


27 Pickel and Rabern (2017: 155) call this ‘structure intrinsicalism’, and advocate it.
28 Nor would it help to suggest that the meaning of ∀xF(x) depends upon the separate meanings of

the two composite expressions ∀x and F(x). For if we think that open formulas possess semantic values
(in isolation), we will obtain an exactly parallel (and exactly as confused) ‘antinomy of the open formula’
as follows: clearly F(x) and F(y) are notational variants, and so should have the same semantic value; but
they cannot have the same value, since F(x) ∧ ¬F(y) is not a contradiction.
22 logics and languages

certain philosophically interesting dimensions. We explore these philosophical is-


sues in later chapters; here, we simply outline its technicalities.
First-order logic can be thought of as allowing quantification into name position.
For example, if φ(c) is a formula containing a constant symbol c, then we also have
a formula ∀vφ(v/c), replacing c with a variable which is bound by the quantifier.
To extend the language, we can allow quantification into relation symbol or function
symbol position. For example, if φ(R) is a formula containing a relation symbol R,
we would want to have a formula ∀Xφ(X/R), replacing the relation symbol R with
a relation-variable, X, which is bound by the quantifier. Equally, if φ( f ) is a formula
containing a function symbol f , we would want to have a formula ∀pφ(p/ f ).
Let us make this precise, starting with the syntax. In addition to all the symbols
of first-order logic, our language adds some new symbols:
• relation-variables: U, V, W, X, Y, Z
• function-variables: p, q
both with numerical subscripts and superscripts as necessary. In more detail: just
like relation symbols and functions symbols, these higher-order variables come
equipped with a number of places, indicated (where helpful) with superscripts. So,
together with the subscripts, this means we have countably many relation-variables
and function-symbols for each number of places. We then expand the recursive
definition of a term, to allow:
• q n (t 1 , …, t n ), for any L -terms t 1 , …, t n and n-place function-variable q n
and we expand the notion of a formula, to allow
• X n (t 1 , …, t n ), for any L -terms t 1 , …, t n and n-place relation-variable X n
• ∃X n φ and ∀X n φ, for any n-place relation-variable X n and any second-order
L -formula φ which contains neither of the expressions ∃X n nor ∀X n
• ∃q n φ and ∀q n φ, for any n-place function-variable q n and any second-order
L -formula φ which contains neither of the expressions ∃q n nor ∀q n
We will also introduce some abbreviations which are particularly helpful in a
second-order context. Where Ξ is any one-place relation symbol or relation-
variable, we write (∀x : Ξ)φ for ∀x(Ξ(x) → φ), and (∃x : Ξ)φ for ∃x(Ξ(x)∧ φ).
We also allow ourselves to bind multiple quantifiers at once; so (∀x, y, z : Ξ)φ ab-
breviates ∀x∀y∀z((Ξ(x) ∧ Ξ(y) ∧ Ξ(z)) → φ).

1.10 Full semantics


The syntax of second-order logic is straightforward. The semantics is more subtle;
for here there are some genuinely non-equivalent options.
We start with full semantics for second-order logic (also known as standard se-
mantics). This uses L -structures, exactly as we defined them in Definition 1.2.
1.10. full semantics 23

The trick is to add new semantic clauses for our second-order quantifiers. In fact,
we can adopt any of the Tarskian, Robinsonian, or Hybrid approaches here, and we
sketch all three (leaving the reader to fill in some obvious details).
Tarskian. Variable-assignments are the key to the Tarskian approach to first-order
logic. So the Tarskian approach to second-order logic must expand the notion of
a variable-assignment, to cover both relation-variables and function-variables. In
particular, we take it that σ is a function which assigns every variable to some entity
a ∈ M, every n-place relation-variable to some subset of M n , and every function-
variable to some function M n Ð→ M. We now add clauses:

M, σ ⊧ X n (t 1 , …, t n ) iff (t M,σ
1 , …, t n ) ∈ (X )
M,σ n M,σ
for any
L -terms t 1 , …, t n
M, σ ⊧ ∀X φ(X ) iff M, τ ⊧ φ(X n ) for every variable-assignment τ
n n

which agrees with σ except perhaps on X n


M, σ ⊧ ∀q n φ(q n ) iff M, τ ⊧ φ(q n ) for every variable-assignment τ
which agrees with σ except perhaps on q n

Robinsonian. The key to the Robinsonian approach to first-order logic is to in-


troduce a new constant symbol for every entity in the domain. So the Robinsonian
approach to second-order logic must introduce a new relation symbol for every pos-
sible relation on M, and a new function symbol for every possible function. Let
M◾ be the structure which expands M in just this way. So, for each n and each

S ⊆ M n , we add a new relation symbol RS with S = RM S , and for◾ each function
g : M n Ð→ M we add a new function symbol f g with g = f M g . We can now
simply rewrite the first-order semantics, replacing M○ with M◾ , and adding:

M◾ ⊧ ∀X n φ(X n ) iff M◾ ⊧ φ(RS /X n ) for every S ⊆ M n


M◾ ⊧ ∀q n φ(q n ) iff M◾ ⊧ φ( f g /q n ) for every function g : M n Ð→ M

Hybrid. The key to the Hybrid approach to second-order logic is to define, up-
front, the three-place relation between M, a formula φ, and a relation (or function)
on M.29 We illustrate the idea for the case of relations (the case of functions is ex-
actly similar). Let S be a relation on M n . Let RS be an n-place relation symbol not
occurring in L . We define M[S] to be a structure whose signature is L together
M[S]
with the new relation symbol RS , such that M[S]’s L -reduct is M and RS = S.
Then where φ(X) is an L -formula with free relation-variable displayed, we define:

M ⊧ φ(S) iff M[S] ⊧ φ(RS /X) for any relation symbol RS ∉ L


M ⊧ ∀X n φ(X n ) iff M ⊧ φ(S) for every relation S ⊆ M n
29 Trueman (2012) recommends a semantics like this as a means for overcoming philosophical resis-
tance to the use of second-order logic.
24 logics and languages

The three approaches ultimately define the same semantic relation. And we call the
ensuing semantics full second-order semantics.
The relative merits of these three approaches are much as before. So: the
Tarskian approach unhelpfully treats relation-variables as if they were varying pred-
icates; the Robinsonian approach forces us to stretch the idea of a language to break-
ing point; but the Hybrid approach avoids both problems and provides us with a
reasonable notion of compositionality. (It is worth noting, though, that all three
approaches effectively assume that we understand notions like ‘all subsets of M n ’.
We revisit this point in Part B.)

1.11 Henkin semantics


The Tarskian, Robinsonian, and Hybrid approaches all yielded the same relation,
⊧. However, there is a genuinely alternative semantics for second-order logic. More-
over, the availability of this alternative is an important theme in Part B of this book.
So we outline that alternative here.
In full second-order logic, universal quantification into relation-position effec-
tively involves considering all possible relations on the structure. Indeed, using
℘(A) for A’s powerset, i.e. {B : B ⊆ A}, we have the following: if X is a one-place
relation-variable, then the relevant ‘domain’ of quantification in ∀Xφ is ℘(M); and
if X is an n-place relation-variable, then the relevant ‘domain’ of quantification in
∀Xφ is ℘(M n ). An alternative semantics naturally arises, then, by considering
more restrictive ‘domains’ of quantification, as follows:

Definition 1.6: A Henkin L -structure, M, consists of:


• a non-empty set, M, which is the underlying domain of M
n ⊆ ℘(M ) for each n < ω
a set M rel n

n ⊆ {g ∈ ℘(M
a set M fun ) : g is a function M n Ð→ M} for each n < ω
n+1

• an object c ∈ M for each constant symbol c from L
M

• a relation RM ⊆ M n for each n-place relation symbol R from L


• a function f M : M n Ð→ M for each n-place function symbol f from L .

In essence, M rel
n serves as the domain of quantification for the n-place relation-
variables, and M fun
n serves as the domain of quantification for the n-place function-
variables. As before, though, we can make this idea precise using any of our three
approaches to formal semantics. We sketch all three.
Tarskian. Where M is a Henkin structure, we take our variable-assignments σ
to be restricted in the following way: σ assigns each variable to some entity a ∈ M,
each n-place relation-variable to some element of M rel
n , and each n-place function-
fun
variable to some element of M n . We then rewrite the clauses for the full semantics,
1.11. henkin semantics 25

exactly as before, but using this more restricted notion of a variable-assignment.


Robinsonian. Where M is a Henkin structure, we let M◽ be the structure which

expands M by adding new relation symbols RS such that S = RM S for every relation

S ∈ M rel
n , and new function symbols f g such that g = f g for every function g ∈
M

M fun
n . We then offer these clauses:

M◽ ⊧ ∀X n φ(X n ) iff M◽ ⊧ φ(RS /X n ) for every relation S ∈ M rel


n
M◽ ⊧ ∀q n φ(q n ) iff M◽ ⊧ φ( f g /q n ) for every function g ∈ M fun
n

Hybrid. We need only tweak the recursion clauses, as follows:

M ⊧ ∀X n φ(X n ) iff M ⊧ φ(S) for every relation S ⊆ M rel


n
M ⊧ ∀q n φ(q n ) iff M ⊧ φ(g) for every function g ∈ M fun
n

We say that Henkin semantics is the semantics yielded by any of these three ap-
proaches, as applied to Henkin structures. Importantly, Henkin semantics gener-
alises the full semantics of §1.10. To show this, let M be an L -structure in the sense
of Definition 1.2. From this, define a Henkin structure N by setting, for each n < ω,
n = ℘(N ) and N n as the set of all functions N Ð→ N. Then full satisfaction,
N rel n fun n

defined over N, is exactly like Henkin satisfaction, defined over N.


The notion of a Henkin structure may, though, be a bit too general. To see
why, consider a Henkin L -structure M, and suppose that R is a one-place rela-
tion symbol of L , so that RM ⊆ M. Presumably, we should want M to satisfy
∃X∀v(R(v) ↔ X(v)), for RM should itself provide a witness to the second-order
existential quantifier. But this holds if and only if RM ∈ M rel
1 , and the definition of
a Henkin structure does not guarantee this. For this reason, it is common to insist
that the following axiom schema should hold in all structures:
Comprehension Schema. ∃X n ∀v(φ(v) ↔ X n (v)), for every formula φ(v) which
does not contain the relation-variable X n
We must block X n from appearing in φ(v), since otherwise an axiom would be
∃X∀v(¬X(v) ↔ X(v)), which will be inconsistent. However, we allow other free
first-order and second-order variables, because this allows us to form new concepts
from old concepts. For instance, given the two-place relation symbol R, we have as
an axiom ∃X 2 ∀v1 ∀v2 (¬R(v1 , v2 ) ↔ X 2 (v1 , v2 )), i.e. M rel
2 must contain the set of
all pairs not in RM , i.e. M 2 ∖ RM . So: if we insist that (all instances) of the Com-
prehension Schema must hold in all Henkin structures, then we are insisting on
further properties concerning our various M rel n s. There is also a predicative version
of Comprehension:
Predicative Comprehension Schema. ∃X n ∀v(φ(v) ↔ X n (v)), for every formula
φ(v) which neither contains the relation-variable X n nor any second-order quantifiers
26 logics and languages

When we want to draw the contrast, we call the (plain vanilla) Comprehension
Schema the Impredicative Comprehension Schema. But this will happen only
rarely; we only mention Predicative Comprehension in §§5.7, 10.2, 10.c, and 11.3.
We could provide a similar schema to govern functions. But it is usual to make
the stronger claim, that the following should hold in all structures (for every n):30
Choice Schema. ∀X n+1 (∀v∃y X n+1 (v, y) → ∃p n ∀v X n+1 (v, p n (v)))
To understand these axioms, let S be a two-place relation on the domain, and sup-
pose that the antecedent is satisfied, i.e. that for any x there is some y such that
S(x, y). The relevant Choice instance then states that there is then a one-place func-
tion, p, which ‘chooses’, for each x, a particular entity p(x) such that S(x, p(x)).
For obvious reasons, this p is known as a choice function. Hence, just like the Com-
prehension Schema, the Choice Schema guarantees that the domains of the higher-
order quantifiers are well populated.
This leads to a final definition: a faithful Henkin structure is a Henkin structure
within which both (impredicative) Comprehension and Choice hold.31

1.12 Consequence
We have defined satisfaction for first-order logic and for both the full- and Henkin-
semantics for second-order logic. However, any definition of satisfaction induces a
notion of consequence, via the following:

Definition 1.7: A theory is a set of sentences in the logic under consideration. Given a
structure M and a theory T, we say that M is a model of T, or more simply M ⊧ T, iff
M ⊧ φ for all sentences φ from T. We say that T has φ as a consequence, or that T
entails φ, or more simply just T ⊧ φ, iff: if M ⊧ T then M ⊧ φ for all structures M.

Note that this definition is relative to a semantics. So there are as many notions of
logical consequence as there are semantics.
Here are some examples to illustrate the notation. Consider the natural numbers
N and the integers Z in the signature consisting just of the symbol <, where this
is given its natural interpretation. It is easy to see that both structures satisfy the
following axioms:
∀x∀y∀z((x < y ∧ y < z) → x < z)
∀x(x ≮ x)
∀x∀y(x < y ∨ x = y ∨ y < x)
30 For more, see Shapiro (1991: 67).
31 See e.g. Shapiro (1991: 98–9).
1.13. definability 27

These are the axioms of a linear order. Let TLO be the theory consisting of just these
three axioms. Then we would write N ⊧ TLO and Z ⊧ TLO . But if we drop the third
axiom, we obtain the related notion of a partial order. For an example of a partial
order which is not a linear order, consider any set X with more than two elements,
and consider the structure P whose first-order domain is the powerset ℘(X) of X,
with < interpreted in P as the subset relation. If a, b are distinct elements of X, then
P ⊧ {a} ≮ {b} ∧ {a} ≠ {b} ∧ {b} ≮ {a}. So P ⊭ TLO .

1.13 Definability
In addition to a notion of consequence, a semantics will induce a notion of defin-
ability, as follows:

Definition 1.8: Let M be any structure and n ≥ 1. We say that a subset X of M n


is definable iff there is both a formula φ(v1 , …, vn , x1 , …, xm ) with all free variables
displayed and also elements b1 , …, bm ∈ M such that:

X = {(a1 , …, an ) ∈ M n : M ⊧ φ(a1 , …, an , b1 , …, bm )}

Here, the elements b1 , …, bm are called parameters. Many authors allow parame-
ters to be tacitly suppressed, and so say that X is definable iff X = {(a1 , …, an ) ∈
M n : M ⊧ φ(a1 , …, an )} for some φ(v1 , …, vn ) which is (tacitly) allowed to con-
tain further unmentioned parameters. If parameters are not allowed, such authors
typically say this explicitly. We will be similarly explicit. When parameters are not
allowed, the resulting sets are called parameter-free definable sets. Clearly a set is M-
definable iff it is parameter-free definable in some signature-expansion of M (see
Definition 1.4).
To illustrate the idea of definability, consider again the natural numbers N in the
signature consisting just of <, again with its natural interpretation. Here is a simple
definable set:
{0} = {n ∈ N : N ⊧ ¬∃x x < n}

As a slightly more complicated example, the graph of the successor operation in N


is definable, since intuitively n = m + 1 iff m is less than n and there is no natural
number strictly between m and n. More precisely:

G = {(n, m) ∈ N 2 : N ⊧ (m < n ∧ ¬∃z(m < z < n))}

Now, both of these sets are parameter-free definable. And so it follows that all defin-
able sets over N are parameter-free definable. For, where S is the successor function
28 logics and languages

on the natural numbers, each natural number n is equal to the term S n (0), which
we define recursively as follows:

S 0 (a) = a S n+1 (a) = S(S n (a)) (numerals)

(We label this definition ‘(numerals)’ for future reference.) Hence, to say that 2 =
S 2 (0) is just a fancy way of saying that two is the second successor of zero. The
terms S n (0) are sometimes called the numerals, and clearly N ⊧ n = S n (0) for
each natural number n ≥ 0. So, we can explicitly define the numerals in terms of
the less-than relation using G, any definable set on N is parameter-free definable, by
the following:

{(a1 , …, an ) ∈ N n : N ⊧ φ(a1 , …, an , b1 , …, bm )}
={(a1 , …, an ) ∈ N n : N ⊧ φ(a1 , …, an , S b1 (0), …, S bm (0))}

For an example of a structure with definable sets which are not parameter-free de-
finable, let L be a countable signature and let M be an uncountable L -structure.
Since there are only countably many L -formulas, there are only countably many
parameter-free definable sets. But trivially the singleton {a} of any element a from
M is definable, as {a} = {x ∈ M : M ⊧ x = a}. So M has uncountably many
definable subsets which are not parameter-free definable.
Finally, it is worth mentioning a particular aspect of definability in second-order
logic. Consider the natural numbers N in the full semantics, and consider the set
{(n, A) ∈ N × ℘(N) : N ⊧ A(n)} consisting of all pairs of numbers and sets
of numbers such that the number is in the set. It obviously makes good sense to
say that this set is definable, even though it is not a subset of N × N but rather of
N × ℘(N). So, in the case of second-order logic, we expand the notion of definabil-
ity to include both subsets of products of the second-order domain, and subsets of
products of the first-order domain and the second-order domain. This point holds
for both the Henkin and the full semantics.

1.a First- and second-order arithmetic


We have laid down the syntax and semantics for the logics which occupy us
throughout this book. However, we will frequently discuss certain specific mathe-
matical theories. So, for ease of reference, in this appendix we lay down the usual
first- and second-order axioms of arithmetic. We cover set theory in the next ap-
pendix, and reserve all philosophical commentary for later chapters.

Definition 1.9: The theory of Robinson Arithmetic, Q, is given by the universal closures
of the following eight axioms:
Another random document with
no related content on Scribd:
testes magyar történeti könyv nem lévén forgalomban. Amilyen
pompás részlet-munkálatok vannak a magyar történet területéről,
amely művek azonban, természetüknél fogva, nem jutnak a
nagyközönség elé, olyan árva az összefoglaló irodalom, amelyben
majdnem kizárólag csak iskolakönyvek, vagy iskolaszerü könyvek
jelennek meg. Egyik rosszabb és korlátoltabb a másiknál. Az
Akadémiának, noha ezt nem mondta, igaza van. Gyatrák e magyar
történetek és semmi egyébre nem alkalmasok, mint arra, hogy az
olvasót félrevezessék. Vezetik is. Nincs a magyar közép-
műveltségnek egyetlen olyan embere sem, aki tisztában volna akár
Nagy Lajos, akár Mátyás uralkodásával. Pedig e két alak nagyon
történeti s a szaktudomány előtt már a leghatározottabb
körvonalakkal áll. E szaktudományt azonban a történeti
kompendiumok irói nem hajlandók tudomásul venni. Példa: a
magyar honfoglalásnak van egy ragyogó történetirója, Sebestyén
Gyula a neve. Valósággal testi gyönyörüség elolvasni, hogy a
rendelkezésre álló gyér és keverék-adatokból ez az ember hogyan
hámozta ki egyrészt a magyar honfoglalásnak a történetét, másrészt
a monda-világát, hogyan választotta el a kettőt egymástól, hogyan
vitte tálcán be a történetbe azt, aminek csak kusza nyomai voltak
elszórva. Az iskolai történetirás egyszerüen nem vett tudomást
Sebestyén Gyula munkálkodásának eredményéről. És nem vett
tudomást Acsády munkálkodásáról, sőt, II. Józsefet illetően,
Marczaliéról sem. Sebestyén Gyula megmondja és bebizonyitja,
hogy ki volt Anonymus, csodaszépen mutatja ki, hogy a neve
Adorján volt, III. Béla kancellistája volt, püspök volt Erdélyben. Ön, jó
olvasóm, azonban még most is ugy tudja, hogy Anonymus
valamelyik Béla király névtelen jegyzője volt. Mert ön igy tanulta az
iskolában, s mert a fiától, aki most tanulja, ugyanigy értesül.
Az ön fia meg éppugy megtanulja a gyász-magyarokról, vagy
gaz-magyarokról szóló csacsiságot, ahogy ön is megtanulta. Lehelt
és Botondot is tanulja, valamint azt is, hogy az Árpád-királyok
énekmondói hegedősök voltak. Mindez nem igaz, nem igy igaz,
sokkal szebb és érdekesebb maga az igazság. De azt még nem
terjesztik. Az Akadémia sem, amely egyszerüen nem adja ki a
pályadijat, de nem néz körül s nem látja meg, hogy van már magyar
történetirás, amelynek kiadhatná.

A PERZSÁK

Az 1914. év Vojnich-jelentésében nem volt semmi különösen


bántó, de hogy a garat mégse járjon üresen, benne volt az az egy
mondat, hogy a termés nem valami kiváló, s ha mindent
összeveszünk, a szinpadi irodalom legfontosabb, vagy
leghasznosabb, vagy leggyönyörübb ténye Ajszkülosz tragédiájának,
a Perzsáknak, a leforditása volt. Hódolat Ajszkülosznak, aki tán mind
a mai napig a legnagyobb drámairó, hódolat a perzsáknak, akik az
ókornak olyan csodálatos népei voltak, hogy mi, akik az ókorból csak
a görögökről és a rómaiakról tudunk (mert a többi adatok elől még az
iskolában bedugják a fülünket), mi egyáltalán nem tudjuk, hogy
milyenek voltak a perzsák s a többi ázsiaiak, akik hatalmas
kulturéletet fejeztek be akkor, amikor a görögök még alighogy
elkezdték. Külön tisztelet a Perzsáknak, a darabnak, amely alkalmat
adott rá, hogy egy magyar iró a hatvanon felül, megmutassa: ősz
ember is milyen ifjonti erővel és lelkesedéssel tud dolgozni, ha a
feladata tüzeli, köszönet tehát Rákosi Jenőnek, hogy a maga hajlott
korában nekiült és a Perzsákat leforditotta s még egy köszönet Ivánfi
Jenőnek, aki a darabot annak idején rendezte s egypár megrázó
tömegjelenetet produkált. Hódolat, hála és köszönet mindenkinek,
akit megillet; amint méltóztatnak látni, nem óhajtok egyetlen
levelecske babért sem elvenni senkitől. De 1914-ben azt mondani,
hogy az 1913. év magyar szinpadi irodalmának a legnagyobb
eseménye Ajszkülosz Perzsáinak a leforditása volt, ez mégis erős
egy kicsit. Az Akadémia joggal becsüli tul a forditást ennyire, hiszen
tagjainak legnagyobb része nem csinált egyebet egész életében,
csak forditott, néha ugy, hogy az eredeti szerző nevét elfeledte
megemliteni: – azonban ilyen vélekedéssel a nyilvánosság elé
kilépni, mégis merészség. A magyar irodalom ma már jár annyira a
maga lábán, – s több évtized óta teszi ezt – hogy egy forditás nem
lehet benne a legfontosabb esemény, ha akárki is a leforditott, sőt
majdnem azt lehet mondani: ha akárki is a forditó. Már Kazinczynak
nem volt igaza, s hogy nincs igaza, már a kortársai érezték, mikor
minden falusi plebánost, aki valamely beteg idegen románt magyarra
forditott, istenfinak koszoruzott. S ez száz esztendővel ezelőtt volt.
Azóta a forditások nyilván sokat veszitettek jelentőségükből, s a
magyar Shakspere sem lett volna az ötvenes éveken tulnan és innen
olyan jelentékeny esemény, ha Petőfi, Vörösmarty és Arany János
nem forditottak volna Shakspereből egynéhány darabot. Arany
János Arisztofanesze, amely 1880-ban jelent meg, irodalmi esemény
volt, mert Arany János adta ajándékba az irodalomnak az egész
Arisztofaneszt.
Mert Arany János – ezen van a hangsuly. Azonban ez is csak
irodalmi esemény volt, a szinpadi irodalomnak kevés köze lehetett
hozzá s nem hiszem, hogy ha akkor lett volna Vojnich-dij, a birálók,
akiknek akkoriban Csiky Gergely, Rákosi Jenő, Dóczy Lajos, Berczik
Árpád darabjairól kellett beszámolniuk (mert akkoriban ezek irták a
darabokat), e nevezett fiatal irókat azzal ütötték volna fejbe, hogy ők
is irtak ugyan, de a szinpadi irodalom legfontosabb eseménye mégis
csak Arany János Arisztofanesze.
Arany János, akinek kinos igazságérzete volt minden-életében,
okvetlenül tiltakozott volna ez ellen, s megmondta volna, hogy
Arisztofanesznek, még ha játszanák is, igen kevés köze van az
1880. év magyar szinpadi irodalmához. Tessék csak elgondolni,
mennyivel kevesebb még, egyetlenegy Ajszkülosz-tragédiának,
1913-ban! Még akkor is, ha Rákosi Jenőnél nagyobb nyelvművész
nem volna a ma élő magyar irodalomban, még akkor is, ha a
Perzsák forditása örök-életü minta-műve volna a magyar nyelven
megszólaló antik metrumoknak.
Erről azonban nincs szó, mert ahogy felnézek a
könyvespolcomra és ott sorakozni látom Rákosi Jenő husz kötetét,
érzem, hogy nekem van igazam, s nem az Akadémiának, amely
Rákosi Jenőből – aki nyilván nem is kéri ezt a dicsőséget –
klasszikust akar csinálni, holott Rákosi Jenőnek a legnagyobb
érdeme az, hogy romantikus mert lenni, amikor ez épp olyan bün
volt, mint amilyen bün ma a modernség. Egyébként nem is igaz,
hogy az Akadémia dicsőséget akart szerezni Rákosi Jenőnek:
dehogy. Az Akadémia senkinek sem akar dicsőséget szerezni, csak
ütni akart egyet a mai magyar szinpadi irodalmon. Az Akadémia még
Ajszküloszt sem akarta megtisztelni, mert ha ezt akarná, akkor
csináltatna egy végleges és el nem avuló teljes Ajszküloszt –
mondjuk például Rákosi Jenővel. De az Akadémia ezt biztosan nem
csináltatja. Nagyon érdekli az Akadémiát Ajszkülosz! Sem
Ajszkülosz, sem Euripidész, sem a többiek, akik pedig már eléggé
öregek ahhoz, hogy az Akadémiát érdekelhetnék. Az Akadémia,
hálaistennek, nem csinál semmit, s ha néha mégis kiduzzad belőle a
tetterő, akkor a kezébe kap egy korosabb irót és azzal lesujt az
ifjabbak közé. Vagyis: sem a korosat, sem az ifjakat nem engedi
dolgozni. Ilyen a magyar tudományos Akadémia.
Hogy mit gondolt Széchényi István gróf (akinek a műveit egy
tisztességes kritikai kiadásban szintén nem tudtuk mindezideig
megkapni az Akadémiától), mikor ezt a műintézetet megalapitotta,
nem tudom. Csak azt tudom, hogy a gőzhajózást is Széchényi
kezdte el Magyarországon, ellenben mikor elkezdte, akkor misem
állt távolabb tőle, mint az a gondolat, hogy 1913-ban ugyanaz a
gőzhajó járjon Budapesttől Zimonyig, amelyiket ő 1831 március 16-
án utnak inditott.
UTAZOM

MÉLÁZÁS

Események… A szememmel eszem fel őket a lapokból és mélán


kérődzöm rajtuk, mint az ökör, mert most szombat van, nem kell
sietni, hiszen holnap vasárnap lesz. Igaz, hogy aztán ujra kezdődik a
hétfő meg a többi hétköznap, de mindegy, azért mégis hat napon át
teremtette az eget és a földet és mindent, de a hetedik napon…
Szombat van mégis, bennem van a ráérek érzése, szombat, az egy
kis őrházmegálló az élet visinálisának másodrendü pályateste
mellett. Ez a hasonlat az élet vicinálisával reggel jutott eszembe,
mikor nagyon kialvatlan és komor voltam, de azért ellenőrizni
tudtam, hogy ez a hasonlat nem uj és mingyárt felötlött az
emlékezetemben az a vidéki lakodalom, ahol ezt a hasonlatot
hallottam. Hogy az élet olyan, mint a vasut, van egy kiinduló pontja,
a születés (keleti pályaudvar, vagy nyugoti), aztán az élet vonatja
szalad, szalad és megáll kisebb-nagyobb állomások mellett. Kisebb
állomások az évvégi vizsgák a gyereknél, nagyobbak az első hosszu
nadrág, vagy hosszu szoknya, az első nemi ösztön és igy tovább,
ahogy parancsolják. A legnagyobb állomás, mondta a szónok: a
házasság.
*
Itt képzavar jött a szónok részéről, mert ezt az állomást révnek
nevezte, melybe a hajók befutnak. Az ő részéről volt a képzavar,
mondom, mert ő nem tudhatta, tudom, hogy nem is tudta, hogy
vannak ilyen révek, amikbe begurul a vonat és besiklik a hajó.
Milyen szépek vannak! Az idén csak Folkstone és Boulogne-sur-Mer
maradt a szememben; egyik szebb volt, mint a másik. Ahogy a
smaragd-zöld angol mezőkön átcsuszik a vonat (mert hihetetlenül
halk járása van), aztán megáll a különösen kék vagy zöld tenger
felett, aztán egyszerre lekattog, le egyenesen a tengerbe, pontosan
a hajó mellé. Ki lehet szállni a vonatból, jó, hogy ki lehet szállni, mert
van velünk egy angol pap, nagyon kedves ember, kis kurta pipa a
szájában, egy Everybody-kötet a kezében és szorgalmasan olvas,
de közben folyton cirpel, mint a tücsök, vagyis a fogát szijja, és
szipákol, ami nem annyit jelent, hogy a pipáját szivja meg, hanem
azt, hogy az orrával minden második másodpercben hangosan
szippantja a levegőt. Nem lehet kibirni, pedig nagyon kedves ember.
*
De miért jut az eszembe ez, mikor nem erről akartam beszélni?
Nem is Boulogne-sur-Merről, ahol megint beuszik a hajó a
világitótornyok közt a kikötőbe és megint olyan jó beülni a francia
vonatba; francia hordárok, lehet velük beszélni, értem a nyelvüket,
az étkezőben hosszu francia kenyér, a szomszéd asztalnál szakállas
francia vidékiek. Azt hiszem, egy pénzügyigazgató, egy
polgármester és a képviselőjük. A képviselő aufsnejdol, nagy
neveket mond, a pénzügyigazgató és a polgármester lesik a száját,
aztán ők beszélnek. Bizonyisten olyanok, mint a magyar vidékiek.
Hangosan beszélnek, inkább orditanak, csemcsegnek, zsiros a
pofájuk, ahogy esznek. Jó nézni őket. Piros bort isznak és mellettünk
csodapiros pipacsos mezők rohannak el. Estére Párisba érünk.
*
Halt, halt, nem erről van szó. Huzigálom a gondolataimat ide-oda,
nagyon szétjárnak. Arról akartam beszélni, hogy a rendőrök
agyonvertek egy embert Szabadkán. Ötszáz pár bocskort lopott
állitólag, vagy tán nem is lopta, mindenesetre nem akarta bevallani,
aztán reggel halva találták! Na igen, van ugy néha, hogy halva
találnak egy embert. Ez a végállomás az élet vicinálisán. Kiszállni,
mondják, tessék kiszálni. Némelyeknek az érelmeszesedés sugja a
fülébe, másoknak a rendőr csizmája rugja a halántékába. Igen… ez
különböző, de lényegében azonegy. Nazim pasa például… hát
persze, Enver bej kiváló ember, de azért ugy látszik, hiába, a török
katonák elfutnak. Nézd, nem akarnak meghalni a hazáért. Mért
nem? Egy latin idézetből emlékszem rá, hogy ez édes és diszes
dolog (dulce et decorum). Ez talán nem volna igaz? Ezt bizonyosan
olyan ember irta, aki nem halt meg a hazáért és egyáltalán nem halt
meg, hanem élt és jól érezte magát. Ahogy vesszük. Az is édes és
diszes dolog.
*
Élni? Irni? Vagy sirni? Meghalni? Most nem tudom, majd
vasárnap kigondolom, akkor ráérek, mert akkor ugy is fáj a fejem.

GYORSÁRU

Utazók nagy pántos ládáit kiskerekü lapos szekérkén dörögve


guritják a perrón végébe. A sineken csenevész vicinális-kocsikat
lökdösnek idébb-odébb, hogy szabad legyen a kijárás. Helyet
csinálnak a személyvonatnak.
– Mingyár itt lesz – mondja egy vasuti. – Kanyarodik már.
Fehér sistergésbe burkoltan siklik el az állomás előtt, megáll, a
kocsik, mintha kifujnák magukat a nagy szaladás után, a tüdejükből
lent gőzt bocsátanak ki. Az ajtók kinyilnak, mindenféle emberek
tottyannak le a magas lépcsőről a nedves földre. Bőrlábszáru vidéki
urak, kurta bundákban, zerge-szakállas kalap a fejükön, tollbóbitás
városi nők, kilenc honvédbaka is, akik mingyár ott a kocsi mellett
kettős rendekbe igazodnak és vezényszóra indulnak el. A postakocsi
a vonat végén van, kétszárnyu ajtaját kitárják; a hivatalnokoknak
nyitva van a bluzuk, nincs gallér az ingükön, gyorsan adják le a
csomagokat, leveleket, gyorsan veszik át őket. Csupa mozgás az
egész vonat-harmonika. Csak a közepén van egy teherkocsi, azt
senkise bántja, fekete kocsi, ni, a közepén egy kis fehér kereszt van
ráirva. Odamegyek hozzá, hogy megnézzem. Kevés cédula van
ráragasztva, hamarjában nem tudom kibetüzni, hogy honnan jön,
csak azt látom, hogy Kolozsvárra megy. Gyorsáru. Aztán egy kis
fehér papirdarabkát veszek észre, nincs semmi hivatalos formája és
egyetlen szó van ráirva. Ez:
Leiche.
Nagy csend támad a fejemben. Holttest. Halott embertársam jön
jég közé csomagolva onnan, ahová én most vissza akarok kerülni.
Halott embertársam siet Kolozsvárra, mint gyorsáru. Na, mit állok itt?
Mi van ezen? Nem halnak-e meg az emberek, csaknem mindnyájan,
engem és azokat kivéve, akiket szeretek? És idegenből haza is kell
utazni, nem? Vagy ugy, hogy beül az ember bársonydiványra,
fölteszi a cekkert a hálóba, betakarja a térdét nagykendővel, a
nagykendőre rátesz egy könyvet, olvas, közben morog, hogy nem jó
a fütés, nézi a könybelábadt ablakokat és népdalokat próbál ráhuzni
a kerékkattogás ritmusára. Vagy ugy, hogy befekszik a koporsóba, a
jég közé, néz, néz, keresztül a kocsi tetején, keresztül az
égboltozaton, keresztül a naprendszeren. Csak fekszik és nem
csinál semmit, nem mutatja a kalauznak, hogy szabadjegy, nem
kérdi meg tőle, hol állunk. Nagyvárad? És hány perc? Öt.
Hol állunk? Sehol. Mindenhol. A jég között. Hány perc? Egy sem.
Ezer év. Mért jöttem el Pestről? És mért megyek vissza? Mért állnak
glédába a katonák és mért sietnek a postatisztek? És az utazók is
mit szaladgálnak össze-vissza a mintaládájukkal. És kinek gyorsáru
ez a halott? Mit akarnak vele? Mért szállitják? Aki meghalt, az mindig
Kolozsváron van. Nincs hideg, de mégis mintha becsuszott volna a
mellembe egy darab jég a szivem fölé. Állok és gondolkodom, hogy
hogy’ utazzam? Van nálam, a nadrágom zsebében egy kis kassza
és benne öt kupos acéljegy a másik utirányba, a jég közé.
Kiváltsam-e valamelyiket? Vagy van még dolgom itt? Futkározni
valóm? Határozni kellene hamar. Hová? A jég közé, vagy a fütött
kocsiba? De kiabálás van. A személyvonat elrándul mellőlem, a
halott továbbutazik. El, el, mi azt hisszük: Kolozsvárra. És jön a pesti
gyors is, már utazom, már eleven vagyok. A sarokban egy finomarcu
nő, gyászkalap alól világit ki az arca fehére. És a lábán barna
antilopszáru fekete lakkcipő, nagyon szép. Már kivánom, már
elindulok Kolozsvárral ellenkező irányban.
Utazom, mint utas. Várjunk, hiszen ráérek; lesz még belőlem
gyorsáru is.
TARTALOM.

Tiz éven alul 3–8


Telefonvicc 9–14
A záróra 15–18
A hizottak 19–23
Pusztul a mozi 24–27
Magyar film 28–31
Mozgó halottak 32–35
Csalnak 36–40
Finis 41–47
Zalán, a barátom 48–52
Dóping 53–59
Apokalipszis 60–64
A 3 idegenek 65–68
A Hortobágy 69–73
A vaspiac 74–80
A börzesztrájk 81–86
Csábitás 87–92
Utazás Horvátországban 93–96
Pallós 97–101
Kéz a zsebben 102–105
A kalauz 106–111
Az átmeneti város 112–116
Helyet, helyet 117–120
Régiségek 121–125
És most világváros? 126–130
Ó, idő! 131–134
Nem fér össze 135–138
Piccadilly 139–143
Meleg étel 144–147
Bazalt 148–154
A kommüniké 155–160
Koránkelők 161–165
Kételkedem 166–170
Főpróba 171–174
A Kegydij 175–180
A fogsor 181–185
Hogyan művelődöm? 186–189
A szerecseny 190–194
Herczeg Ferenc: miniszter 195–199
A Pógár 200–203
Irodalom és kaszinó 204–208
Csipogás 209–214
A goromba külföld 215–218
Magyar népdal 219–224
Önéletrajz 225–230
Felolvasás 231–236
A pályadij 237–241
A Perzsák 242–246
Mélázás 247–250
Gyorsáru 251–254

Nap nyomda részvénytársaság


*** END OF THE PROJECT GUTENBERG EBOOK MIT ÜLTÖK A
KÁVÉHÁZBAN? ***

Updated editions will replace the previous one—the old editions will
be renamed.

Creating the works from print editions not protected by U.S.


copyright law means that no one owns a United States copyright in
these works, so the Foundation (and you!) can copy and distribute it
in the United States without permission and without paying copyright
royalties. Special rules, set forth in the General Terms of Use part of
this license, apply to copying and distributing Project Gutenberg™
electronic works to protect the PROJECT GUTENBERG™ concept
and trademark. Project Gutenberg is a registered trademark, and
may not be used if you charge for an eBook, except by following the
terms of the trademark license, including paying royalties for use of
the Project Gutenberg trademark. If you do not charge anything for
copies of this eBook, complying with the trademark license is very
easy. You may use this eBook for nearly any purpose such as
creation of derivative works, reports, performances and research.
Project Gutenberg eBooks may be modified and printed and given
away—you may do practically ANYTHING in the United States with
eBooks not protected by U.S. copyright law. Redistribution is subject
to the trademark license, especially commercial redistribution.

START: FULL LICENSE


THE FULL PROJECT GUTENBERG LICENSE
PLEASE READ THIS BEFORE YOU DISTRIBUTE OR USE THIS WORK

To protect the Project Gutenberg™ mission of promoting the free


distribution of electronic works, by using or distributing this work (or
any other work associated in any way with the phrase “Project
Gutenberg”), you agree to comply with all the terms of the Full
Project Gutenberg™ License available with this file or online at
www.gutenberg.org/license.

Section 1. General Terms of Use and


Redistributing Project Gutenberg™
electronic works
1.A. By reading or using any part of this Project Gutenberg™
electronic work, you indicate that you have read, understand, agree
to and accept all the terms of this license and intellectual property
(trademark/copyright) agreement. If you do not agree to abide by all
the terms of this agreement, you must cease using and return or
destroy all copies of Project Gutenberg™ electronic works in your
possession. If you paid a fee for obtaining a copy of or access to a
Project Gutenberg™ electronic work and you do not agree to be
bound by the terms of this agreement, you may obtain a refund from
the person or entity to whom you paid the fee as set forth in
paragraph 1.E.8.

1.B. “Project Gutenberg” is a registered trademark. It may only be


used on or associated in any way with an electronic work by people
who agree to be bound by the terms of this agreement. There are a
few things that you can do with most Project Gutenberg™ electronic
works even without complying with the full terms of this agreement.
See paragraph 1.C below. There are a lot of things you can do with
Project Gutenberg™ electronic works if you follow the terms of this
agreement and help preserve free future access to Project
Gutenberg™ electronic works. See paragraph 1.E below.
1.C. The Project Gutenberg Literary Archive Foundation (“the
Foundation” or PGLAF), owns a compilation copyright in the
collection of Project Gutenberg™ electronic works. Nearly all the
individual works in the collection are in the public domain in the
United States. If an individual work is unprotected by copyright law in
the United States and you are located in the United States, we do
not claim a right to prevent you from copying, distributing,
performing, displaying or creating derivative works based on the
work as long as all references to Project Gutenberg are removed. Of
course, we hope that you will support the Project Gutenberg™
mission of promoting free access to electronic works by freely
sharing Project Gutenberg™ works in compliance with the terms of
this agreement for keeping the Project Gutenberg™ name
associated with the work. You can easily comply with the terms of
this agreement by keeping this work in the same format with its
attached full Project Gutenberg™ License when you share it without
charge with others.

1.D. The copyright laws of the place where you are located also
govern what you can do with this work. Copyright laws in most
countries are in a constant state of change. If you are outside the
United States, check the laws of your country in addition to the terms
of this agreement before downloading, copying, displaying,
performing, distributing or creating derivative works based on this
work or any other Project Gutenberg™ work. The Foundation makes
no representations concerning the copyright status of any work in
any country other than the United States.

1.E. Unless you have removed all references to Project Gutenberg:

1.E.1. The following sentence, with active links to, or other


immediate access to, the full Project Gutenberg™ License must
appear prominently whenever any copy of a Project Gutenberg™
work (any work on which the phrase “Project Gutenberg” appears, or
with which the phrase “Project Gutenberg” is associated) is
accessed, displayed, performed, viewed, copied or distributed:
This eBook is for the use of anyone anywhere in the United
States and most other parts of the world at no cost and with
almost no restrictions whatsoever. You may copy it, give it away
or re-use it under the terms of the Project Gutenberg License
included with this eBook or online at www.gutenberg.org. If you
are not located in the United States, you will have to check the
laws of the country where you are located before using this
eBook.

1.E.2. If an individual Project Gutenberg™ electronic work is derived


from texts not protected by U.S. copyright law (does not contain a
notice indicating that it is posted with permission of the copyright
holder), the work can be copied and distributed to anyone in the
United States without paying any fees or charges. If you are
redistributing or providing access to a work with the phrase “Project
Gutenberg” associated with or appearing on the work, you must
comply either with the requirements of paragraphs 1.E.1 through
1.E.7 or obtain permission for the use of the work and the Project
Gutenberg™ trademark as set forth in paragraphs 1.E.8 or 1.E.9.

1.E.3. If an individual Project Gutenberg™ electronic work is posted


with the permission of the copyright holder, your use and distribution
must comply with both paragraphs 1.E.1 through 1.E.7 and any
additional terms imposed by the copyright holder. Additional terms
will be linked to the Project Gutenberg™ License for all works posted
with the permission of the copyright holder found at the beginning of
this work.

1.E.4. Do not unlink or detach or remove the full Project


Gutenberg™ License terms from this work, or any files containing a
part of this work or any other work associated with Project
Gutenberg™.

1.E.5. Do not copy, display, perform, distribute or redistribute this


electronic work, or any part of this electronic work, without
prominently displaying the sentence set forth in paragraph 1.E.1 with
active links or immediate access to the full terms of the Project
Gutenberg™ License.
1.E.6. You may convert to and distribute this work in any binary,
compressed, marked up, nonproprietary or proprietary form,
including any word processing or hypertext form. However, if you
provide access to or distribute copies of a Project Gutenberg™ work
in a format other than “Plain Vanilla ASCII” or other format used in
the official version posted on the official Project Gutenberg™ website
(www.gutenberg.org), you must, at no additional cost, fee or expense
to the user, provide a copy, a means of exporting a copy, or a means
of obtaining a copy upon request, of the work in its original “Plain
Vanilla ASCII” or other form. Any alternate format must include the
full Project Gutenberg™ License as specified in paragraph 1.E.1.

1.E.7. Do not charge a fee for access to, viewing, displaying,


performing, copying or distributing any Project Gutenberg™ works
unless you comply with paragraph 1.E.8 or 1.E.9.

1.E.8. You may charge a reasonable fee for copies of or providing


access to or distributing Project Gutenberg™ electronic works
provided that:

• You pay a royalty fee of 20% of the gross profits you derive from
the use of Project Gutenberg™ works calculated using the
method you already use to calculate your applicable taxes. The
fee is owed to the owner of the Project Gutenberg™ trademark,
but he has agreed to donate royalties under this paragraph to
the Project Gutenberg Literary Archive Foundation. Royalty
payments must be paid within 60 days following each date on
which you prepare (or are legally required to prepare) your
periodic tax returns. Royalty payments should be clearly marked
as such and sent to the Project Gutenberg Literary Archive
Foundation at the address specified in Section 4, “Information
about donations to the Project Gutenberg Literary Archive
Foundation.”

• You provide a full refund of any money paid by a user who


notifies you in writing (or by e-mail) within 30 days of receipt that
s/he does not agree to the terms of the full Project Gutenberg™
License. You must require such a user to return or destroy all
copies of the works possessed in a physical medium and
discontinue all use of and all access to other copies of Project
Gutenberg™ works.

• You provide, in accordance with paragraph 1.F.3, a full refund of


any money paid for a work or a replacement copy, if a defect in
the electronic work is discovered and reported to you within 90
days of receipt of the work.

• You comply with all other terms of this agreement for free
distribution of Project Gutenberg™ works.

1.E.9. If you wish to charge a fee or distribute a Project Gutenberg™


electronic work or group of works on different terms than are set
forth in this agreement, you must obtain permission in writing from
the Project Gutenberg Literary Archive Foundation, the manager of
the Project Gutenberg™ trademark. Contact the Foundation as set
forth in Section 3 below.

1.F.

1.F.1. Project Gutenberg volunteers and employees expend


considerable effort to identify, do copyright research on, transcribe
and proofread works not protected by U.S. copyright law in creating
the Project Gutenberg™ collection. Despite these efforts, Project
Gutenberg™ electronic works, and the medium on which they may
be stored, may contain “Defects,” such as, but not limited to,
incomplete, inaccurate or corrupt data, transcription errors, a
copyright or other intellectual property infringement, a defective or
damaged disk or other medium, a computer virus, or computer
codes that damage or cannot be read by your equipment.

1.F.2. LIMITED WARRANTY, DISCLAIMER OF DAMAGES - Except


for the “Right of Replacement or Refund” described in paragraph
1.F.3, the Project Gutenberg Literary Archive Foundation, the owner
of the Project Gutenberg™ trademark, and any other party
distributing a Project Gutenberg™ electronic work under this
agreement, disclaim all liability to you for damages, costs and
expenses, including legal fees. YOU AGREE THAT YOU HAVE NO
REMEDIES FOR NEGLIGENCE, STRICT LIABILITY, BREACH OF
WARRANTY OR BREACH OF CONTRACT EXCEPT THOSE
PROVIDED IN PARAGRAPH 1.F.3. YOU AGREE THAT THE
FOUNDATION, THE TRADEMARK OWNER, AND ANY
DISTRIBUTOR UNDER THIS AGREEMENT WILL NOT BE LIABLE
TO YOU FOR ACTUAL, DIRECT, INDIRECT, CONSEQUENTIAL,
PUNITIVE OR INCIDENTAL DAMAGES EVEN IF YOU GIVE
NOTICE OF THE POSSIBILITY OF SUCH DAMAGE.

1.F.3. LIMITED RIGHT OF REPLACEMENT OR REFUND - If you


discover a defect in this electronic work within 90 days of receiving it,
you can receive a refund of the money (if any) you paid for it by
sending a written explanation to the person you received the work
from. If you received the work on a physical medium, you must
return the medium with your written explanation. The person or entity
that provided you with the defective work may elect to provide a
replacement copy in lieu of a refund. If you received the work
electronically, the person or entity providing it to you may choose to
give you a second opportunity to receive the work electronically in
lieu of a refund. If the second copy is also defective, you may
demand a refund in writing without further opportunities to fix the
problem.

1.F.4. Except for the limited right of replacement or refund set forth in
paragraph 1.F.3, this work is provided to you ‘AS-IS’, WITH NO
OTHER WARRANTIES OF ANY KIND, EXPRESS OR IMPLIED,
INCLUDING BUT NOT LIMITED TO WARRANTIES OF
MERCHANTABILITY OR FITNESS FOR ANY PURPOSE.

1.F.5. Some states do not allow disclaimers of certain implied


warranties or the exclusion or limitation of certain types of damages.
If any disclaimer or limitation set forth in this agreement violates the
law of the state applicable to this agreement, the agreement shall be
interpreted to make the maximum disclaimer or limitation permitted
by the applicable state law. The invalidity or unenforceability of any
provision of this agreement shall not void the remaining provisions.
1.F.6. INDEMNITY - You agree to indemnify and hold the
Foundation, the trademark owner, any agent or employee of the
Foundation, anyone providing copies of Project Gutenberg™
electronic works in accordance with this agreement, and any
volunteers associated with the production, promotion and distribution
of Project Gutenberg™ electronic works, harmless from all liability,
costs and expenses, including legal fees, that arise directly or
indirectly from any of the following which you do or cause to occur:
(a) distribution of this or any Project Gutenberg™ work, (b)
alteration, modification, or additions or deletions to any Project
Gutenberg™ work, and (c) any Defect you cause.

Section 2. Information about the Mission of


Project Gutenberg™
Project Gutenberg™ is synonymous with the free distribution of
electronic works in formats readable by the widest variety of
computers including obsolete, old, middle-aged and new computers.
It exists because of the efforts of hundreds of volunteers and
donations from people in all walks of life.

Volunteers and financial support to provide volunteers with the


assistance they need are critical to reaching Project Gutenberg™’s
goals and ensuring that the Project Gutenberg™ collection will
remain freely available for generations to come. In 2001, the Project
Gutenberg Literary Archive Foundation was created to provide a
secure and permanent future for Project Gutenberg™ and future
generations. To learn more about the Project Gutenberg Literary
Archive Foundation and how your efforts and donations can help,
see Sections 3 and 4 and the Foundation information page at
www.gutenberg.org.

Section 3. Information about the Project


Gutenberg Literary Archive Foundation

You might also like